You are on page 1of 19

Downloaded from SAE International by University of Michigan, Saturday, July 28, 2018

SAE TECHNICAL
PAPER SERIES 2000-01-0332

A Four-Stroke Homogeneous Charge


Compression Ignition Engine Simulation for
Combustion and Performance Studies
Scott B. Fiveland and Dennis N. Assanis
W. E. Lay Automotive Laboratory, University of Michigan

Reprinted From: Compression Ignition Combustion Processes


(SP–1530)

SAE 2000 World Congress


Detroit, Michigan
March 6-9, 2000

400 Commonwealth Drive, Warrendale, PA 15096-0001 U.S.A. Tel: (724) 776-4841 Fax: (724) 776-5760
Downloaded from SAE International by University of Michigan, Saturday, July 28, 2018

The appearance of this ISSN code at the bottom of this page indicates SAE’s consent that copies of the
paper may be made for personal or internal use of specific clients. This consent is given on the condition,
however, that the copier pay a $7.00 per article copy fee through the Copyright Clearance Center, Inc.
Operations Center, 222 Rosewood Drive, Danvers, MA 01923 for copying beyond that permitted by Sec-
tions 107 or 108 of the U.S. Copyright Law. This consent does not extend to other kinds of copying such as
copying for general distribution, for advertising or promotional purposes, for creating new collective works,
or for resale.

SAE routinely stocks printed papers for a period of three years following date of publication. Direct your
orders to SAE Customer Sales and Satisfaction Department.

Quantity reprint rates can be obtained from the Customer Sales and Satisfaction Department.

To request permission to reprint a technical paper or permission to use copyrighted SAE publications in
other works, contact the SAE Publications Group.

All SAE papers, standards, and selected


books are abstracted and indexed in the
Global Mobility Database

No part of this publication may be reproduced in any form, in an electronic retrieval system or otherwise, without the prior written
permission of the publisher.

ISSN 0148-7191
Copyright © 2000 Society of Automotive Engineers, Inc.

Positions and opinions advanced in this paper are those of the author(s) and not necessarily those of SAE. The author is solely
responsible for the content of the paper. A process is available by which discussions will be printed with the paper if it is published in
SAE Transactions. For permission to publish this paper in full or in part, contact the SAE Publications Group.

Persons wishing to submit papers to be considered for presentation or publication through SAE should send the manuscript or a 300
word abstract of a proposed manuscript to: Secretary, Engineering Meetings Board, SAE.

Printed in USA
Downloaded from SAE International by University of Michigan, Saturday, July 28, 2018

2000-01-0332

A Four-Stroke Homogeneous Charge Compression Ignition


Engine Simulation for Combustion and Performance Studies
Scott B. Fiveland and Dennis N. Assanis
W. E. Lay Automotive Laboratory, University of Michigan

Copyright © 2000 Society of Automotive Engineers, Inc.

ABSTRACT Injection Combustion), and PCIC (Premixed


Compression Ignited Combustion). Engine types have
A computer simulation of the Homogenous Charge ranged from two-stroke to four stroke configurations with
Compression Ignition (HCCI) four-stroke engine has a variety of fuels such as diesel, gasoline, methanol,
been developed for combustion and performance studies. natural gas, and hydrogen. The HCCI process essentially
The simulation couples models for mass, species, and involves a premixed fuel/air mixture that is inducted into
energy within a zero-dimensional framework. The the cylinder at equivalence ratios that can vary from lean
combustion process is described via a user-defined to stoichiometric [6, 7]. Once within the cylinder, the
chemical kinetic mechanism. The CHEMKIN libraries homogeneous fuel/air charge is then compressed until
have been used to formulate a stiff chemical kinetic ignition commences. Ignition leads to a very rapid
solver suitable for integration within a complete engine combustion phase where all heat is released
cycle simulation, featuring models of gas exchange, approximately in 10-35°.
turbulence and wall heat transfer. For illustration, two
The premixed charge, compression ignition HCCI engine
chemical kinetics schemes describing hydrogen and
concept promises to combine the advantages of both the
natural gas chemistry have been implemented in the
Direct Injection, Compression Ignition (DICI) engine and
code. The hydrogen scheme is a reduced one,
the premixed charge, Spark-Ignited (SI) engine, while
consisting of 11 species and 23 reactions. The natural
eliminating their drawbacks. One benefit that can be
gas chemistry is described via the GRI-mechanism 3.0
gained over the current heterogeneous DICI engines is
that considers 53 species and 325 reactions, including
the elimination of the fuel rich zones that are directly
NOx chemistry. Computations are first carried out in a
responsible for pollutant formation, especially
variable volume bomb to demonstrate variations in
particulates [7]. In addition, the homogeneous lean burn
ignition with temperature, pressure, equivalence ratio,
operation will yield lower gas temperatures and hence
and composition. Subsequently, the complete cycle
NOx, as compared to both SI and DICI counterparts.
simulation is exercised to demonstrate the variation in
Furthermore, unthrottled part load operation eliminates
output parameters to charge inlet temperature and
pumping losses leading to improved fuel economy over
effective compression ratio. Overall, this study
the SI engine [1]. In addition, compression ignition
demonstrates the importance of coupling detailed
eliminates SI knock associated with autoignition of fuel-
chemistry descriptions with physical models of the HCCI
air mixture in the end zone, ahead of the advancing flame
engine processes.
front. As the HCCI concept operates on the premise of
autoignition, this allows the use of elevated compression
INTRODUCTION ratios (approximately 20-25:1), unlike in SI engines. The
combination of lean burning, which is thermodynamically
Homogeneous Charge Compression Ignition (HCCI) is attractive and reduces heat transfer losses, and high
currently under widespread investigation due to its compression and thus expansion ratios contributes to
potential to lower NOx and particulate emissions while indicated thermal efficiencies that can approach 55%.
maintaining high thermal efficiency [1, 2, 3, 4, 5]. Residual gas fraction is also reduced with HCCI
Throughout the years it has endured many names in the operation, thus improving volumetric efficiency,
literature: ATAC (Active Thermo-Atmosphere combustion and performance [5]. Overall, HCCI promises
Combustion), LHC (Lean Homogeneous Combustion), to deliver both high thermal efficiencies and reduced
CIHC (Compression Ignited Homogeneous Charge emissions.
Combustion), AR (Active Radical Combustion), HCDC
(Homogeneous Charge Compression Ignition Diesel Although recent investigations into HCCI combustion
Combustion, Diesel Fumigation, MULDIC (Multiple appear promising [4, 5, 6, 7, 8, 9, 10, 11, 12] several
Staged Diesel Combustion, PREDIC (Premixed Direct- problems with the HCCI combustion concept reappear

1
Downloaded from SAE International by University of Michigan, Saturday, July 28, 2018

throughout the literature. Most stem from the fact that the to a system with well understood chemistry. The
HCCI concept gives up two combustion control aspects. sensitivity of the combustion submodel to changes in
First, the timing of ignition is not controlled, neither pressure, temperature, equivalence ratio, and natural gas
indirectly by fuel injection as in a DICI engine, nor directly composition will be demonstrated first in a fundamental,
by the spark as in an SI engine. Second, the rate of heat adiabatic reactor. The behavior of the each of the
release is also not controlled neither by the rate of fuel submodels will then be detailed over the engine cycle,
injection as in a DICI engine, nor by finite turbulent flame and contrasted with other treatments of cyclic processes
propagation as in an SI engine. As a result, the near reported in the literature. Subsequently, parametric
constant volume combustion event leads to a very rapid studies of ignition and engine performance for varying
rate of heat release, thus promoting high mechanical manifold temperature, pressure, and geometric
stresses [8]. In particular, controlling the ignition event at compression are performed for both hydrogen and
higher loads is a widely noted problem. natural gas using the full cycle simulation.
In an effort to understand how mixture preparation and
in-cylinder thermodynamic conditions affect the chemical
MODELING ASSUMPTIONS
kinetics, models of varying resolution have been
developed [4, 7, 9, 10, 11, 12]. These modeling efforts The compression ignition engine simulation of Assanis
provide a good basis for exploring the HCCI combustion and Heywood [15] has provided a solid zero-dimensional
framework for formulation and implementation of the
phenomena. However, in published zero-dimensional
governing equations for the HCCI model. The simulation
simulations, detailed chemistry has typically been
is currently written in a single cylinder version, primarily
coupled to models of turbulence, heat transfer, and gas
because fundamental studies lend themselves to this
exchange that are not predictive over the full engine
configuration. Thermodynamic properties are assumed
cycle. Heat transfer models have been based on data
uniform throughout the chamber volume. The engine
not relevant to HCCI operation, and in many cases the
simulation is a sequence of four-stroke processes. The
gas exchange process has been neglected by specifying
gas exchange process is governed by quasi-steady, one-
conditions at a specific point in the engine cycle. For
dimensional flow equations that are used to predict flow
instance, Smith et al. [7], Van Blarigan and Goldborough
past valves. The compression event is defined from
[5], as well as Kusaka et. al. [12] used the Woshni heat
Intake Valve Closing (IVC) to a transition point prescribed
transfer correlation [13] to predict heat transfer. The
when chemical reactions become important. The
Woschni model [13] was developed from a regression of
combustion event for the HCCI simulation differs from
direct injection diesel engine data and is essentially a
those of the SI and DICI types. As a result of its premixed
linearized radiative/convective relation that will greatly
nature and compression ignition principle, the rate of
over-predict heat transfer in a lean burn, premixed, non-
combustion is strictly limited by the chemical kinetics. In
sooting engine. Furthermore, studies by Poulos and
the reaction regime, the combustion will follow that of
Heywood [14] as well as Assanis and Heywood [15] have
finite rate kinetics and its heat release will be governed
shown that, since turbulence intensity has a large cyclic
by detailed chemistry. Hence, the combustion event has
variation, heat transfer correlations based on multiples of
been modeled using the Chemkin libraries [16], adapted
mean piston speed do not provide a truly predictive
for a variable volume plenum and accounting for heat
capability. In addition, the computational work of Smith et.
transfer effects. The evolution of heat release and
al. [7] was only carried out over the closed part of the
species is governed by a user-defined kinetic scheme.
cycle, starting with bottom dead center of the intake
Later in expansion, the reacting flow mixture will attain
stroke, with no gas exchange. Studies of closed cycle
chemical equilibrium in response to the changing in-
processes are useful for evaluating chemistry, but do not
cylinder conditions, and eventually the composition will
capture intake and exhaust jet flows. Consequently,
become frozen. At this point the mixture is again non-
velocity and length scales of in-cylinder flow structures
reacting and the chemical energy source term tends to
are not accurately represented; thus, wall heat transfer
zero.
losses and pre-ignition boundary conditions can be
inaccurate.
CONSERVATION EQUATIONS
The objective of this work is to develop a full cycle
simulation model of the HCCI engine that would integrate To define a thermodynamic state within the cylinder, two
complex chemistry with physical models of the in-cylinder independent properties and the mixture composition
processes. Emphasis will be placed on integrating need to be known. Consequently, our formulation will
flexible chemical kinetic libraries with models of track the evolution of mass, species, temperature, and
turbulence-based heat transfer and gas exchange pressure throughout the engine cycle. The general
processes for a four-stroke cycle. The paper is arranged equations governing mass, species, and energy will be
as follows. The HCCI model formulation will be presented developed for a variable volume reactor, as shown in
first. This will include a brief development of the Figure 1. It is assumed that the working fluid behaves as
governing equations and various engine submodels (i.e. an ideal gas. The heat transfer is governed by turbulent
gas exchange, combustion, heat transfer) as they apply pipe flow, but derives its characteristic velocity as a

2
Downloaded from SAE International by University of Michigan, Saturday, July 28, 2018

function of mean piston speed, mean velocity, and CONSERVATION OF ENERGY – The generalized
turbulent intensity, which is found from a zero- energy equation for an open thermodynamic system may
dimensional k-ε formulation. be written as:

d (mu ) dV dQ ht
= −p + + ∑m! jh j
dt %$dt# dt#
%$ j
%$# %
"$" #
Internal Displacement Heat Enthalpy
Energy Work Transfer Flux
(5)
Rewriting the first law equation in terms of rate of change
of enthalpy yields,

mh! = ∑ m! j h j − Q! w + P! V − m! h
j
(6)
At this point in the development, an expression needs to
be developed for h! that relates it to the change in mixture
temperature. Assuming a single phase, multi-component
mixture of ideal gases, its enthalpy, h, is defined as
Figure 1. Energy Transfers Associated with Cylinder
Control Volume. h = ∑ Y jh j
j
CONSERVATION OF MASS – The rate of change of (7)
mass within any open system is the net flux of mass and h = h (T , P , Yi ) (8)
across the system boundaries.
where the subcript ‘j’ refers to each of the component
species present in the mixture. Taking partial differentials
dm
= ∑m! j with respect to pressure, temperature and composition:
dt j
(1)
dh ∂h dT ∂h dp ∂h dYi
= + +∑
CONSERVATION OF SPECIES – Equations tracking the dt ∂T P ,Yi dt ∂p T ,Yi
dt ∂
i Yi P ,T,Yi ≠ j
dt
evolution of species within the combustion chamber will (9)
be developed on a mass basis corresponding to the where the partial change of enthalpy with respect to
definition in Eq. (2), pressure at constant temperature and composition is
zero, but pressure effects are accounted for in
mi
Yi = determining changes in composition with allowance for
m (2) dissociation. Hence, Eq. (9) becomes
where ‘m’ denotes the total mass within the control
dh dT dY j dT dY j
cylinder. The species equations are deduced from their = ∑ Y jC p j +∑hj = Cp +∑hj
multi-dimensional counterparts by neglecting species dt j dt i dt dt i dt
(10)
diffusion terms, consistent with the zero-dimensional
assumption. To relate the pressure gradient to the temperature
change, the equation of state is used in its differential
d(mYi ) form. By manipulating the thermodynamic equation of
= ∑m! jYij + S! gen
dt j state for an ideal gas, we get:

R! ! P! m
V ! T!
S! gen = Ωi Wmw v = + − −
(3) R V P m T (11)
Expanding Eq. (3), and applying the continuity equation An expression can now be inserted for the rate of change
yields, of the gas constant R by considering its thermodynamic
dependence, i.e.:
Yi
m
! = ∑ j

!
(
 j
 i i
Ω W
ρ
)
 Y − Y cyl + i mw
(
R = ∑ Y jR j )
j  m 
(4) j
(12)
which is the final form of the species conservation
Taking partial differentials with respect to pressure,
equation.
temperature and composition,

3
Downloaded from SAE International by University of Michigan, Saturday, July 28, 2018

event, it was necessary to describe the evolution of heat


 
dR  ∂R ! 
release via a suitable chemical kinetic mechanism, and a
= ∑  Y j R! j + ∑ Y j well-matched, stiff chemical kinetic solver. CHEMKIN
dt j  ∂
j Yj 
 P ,T, Yi ≠ j  (13)
was selected due to the fact that it is a widely accepted
and used kinetic solver by many researchers in a range
The first term on the R.H.S. of Eq. (13) can be of combustion studies [10, 19]. Simulation development
recognized as the partial expansion of the gas constant has been done in Fortran 90 within the Visual
for a pure substance. Under the problem assumptions, Environment on a PC 333 Mhz Pentium II. A source
this term can be shown to be identically equal to zero, program for a variable volume reactor was written to drive
which is consistent with references [16] and [17]. the CHEMKIN libraries. The driver model includes the
effects of heat transfer, as well as options for different
R! = ∑ R j Y
!
j engine configurations. The program can also operate in
j
(14) an adiabatic standalone mode for fundamental studies.

Using Eqs. (11) and (14), we can obtain an equation that The rates of creation/destruction of chemical species are
relates the rate of change of pressure to the rate of modeled using mass-action kinetics, where the specific
change of temperature, i.e.: reaction rate constants exhibit a strong temperature
dependence. An elementary reaction that involves ‘K’
∑ RiY
!  chemical species in ‘I’ reactions can be represented in
i

!P = P i ! T! V
m ! the form,
 + + − 
 R m T V K K
  (15) ∑ υ 'ki χ k ⇔ ∑ υ 'ki' χ k
k =1 k =1 (19)
Susbstituing Eq. (15) into Eq. (6) yields an equation for
the rate of change of temperature for an ideal gas, I
reacting mixture: Ω k = ∑ (υ''ki − υ' ki )q i
i =1 (20)
  vP∑ R i   υ υ
1  
' ''
1   !
! −m  V !
 K ki K ki
T! = −  ∑ h i − i
Y B + ∑ j j
!
m h + PV  −
 V
 + !
Q w  q i = k fi ∏ [X k ] − k ri ∏ [X k ]
m j
i
A  i R  m  
 
    k =1 k =1 (21)
(16)
The specific reaction rate constant, ‘k’ follow the
Pv
where A = C p − and B = h − Pv (17) Arrhenius dependence where,
T

 − Ei 
MODELING OF RECIPROCATOR PROCESSES k fi = A i T βi exp 
 RT  (22)
GAS EXCHANGE – The one-dimensional quasi-steady
flow model is used to model flow through both the intake THERMODYNAMIC PROPERTY TREATMENT – The
and exhaust valves during the gas exchange processes. fluid in the cylinder is constantly undergoing a change in
Equation (18) is a function of discharge coefficient, valve mixture composition. Once the composition is
area, gas properties, and pressure differential across determined, the partial mixture properties can be
each orifice. The values for both valve lift and discharge summed and appropriately weighted in accordance with
coefficients can be specified or predicted. The values for their mass or mole fractions. This method very simply
cylinder pressure are updated by solving the system of allows thermodynamic property calculation for mixtures
state differential equations through the cycle [18]. containing residual gas, exhaust gas recirculation,
unburned gaseous fuel, air etc. The thermodynamic
 1  property treatment will employ the NASA curve-fits for
 1 1   γ −1  2 
  P o2  2  PT  γ  2γ   PT  γ   specific heat, enthalpy, and entropy:
=  C d A eff      
! i,e
m
 R u To W   1 −  
    Po  γ −1   Po   
  



 C opk 5
(18)
= ∑ a nk T n −1
R n =1 (23)
COMBUSTION – The combustion process in a where the constant of integration a n+1,k R is the enthalpy
homogenous charge compression ignition engine of formation at 0 K.
exhibits no fundamental mixing or entrainment, which
normally control the combustion event for direct-injection h ok 5 a n ,k a
as well as spark ignition engines. As a result the rate of = a 1k + ∑ Tkn −1 + 6 k
RTk n =1 n Tk
heat release is solely driven by the chemical kinetic (24)
reaction rates. To appropriately model the combustion

4
Downloaded from SAE International by University of Michigan, Saturday, July 28, 2018

s ok 4 a n +1, k V
= a 1k ln Tk + ∑ Tkn + a 7 k '=
R n =1 n
(25)
Ab
(31)

HEAT TRANSFER MODEL – The heat transfer model Rewriting Eq. (30) yields
that is utilized was described by Assanis and Heywood 1
[15]. This model is a subset of the k-ε model under the 2C µ C β K  k  2
isotropic turbulence assumption. It uses an energy P=  
cascade between the mean kinetic energy and the 2
3
2 ' m
turbulent kinetic energy. The eddy dissipation is modeled  
3 (32)
via dimensional argument. This model accounts for the
effects of mean piston speed, mean charge motion, and Rapid distortion theory (RDT) is employed to account for
turbulence intensity on the heat transfer coefficient. rapid changes in density. This amplification source term
The definitions for both mean and turbulent kinetic energy was added to the transport equation for turbulent kinetic
are given by Eq. (26). energy

1 du '
K= mU 2 A = 3mu '
2 dt (33)
3
k = mu '2 Since RDT assumes negligible dissipation of the
2 (26) turbulent kinetic energy, conservation of mass and
angular momentum of each eddy is assumed [15, 20].
The equations governing the cascade model are shown
in Eqs. (27)-(29). It is noted that these equations take on 1 1
the classical form for transport phenomena, having an ' ρ  3 u'  ρ  3
=  o  : = 
unsteady term, a convective term, and a source/sink 'o  ρ  u o '  ρ o 
term. The mean flow kinetic energy equation describes (34)
the history of the initial kinetic energy. Over time, this
du '
u dρ '
mean kinetic energy is dissipated to large scale =
turbulence, whose evolution is described by the turbulent dt 3ρ dt
(35)
kinetic energy equation. Finally the eddy dissipation, 2 ρ!
which occurs on a molecular level, dissipates the A= k
3 ρ
turbulent kinetic energy. It is modeled via dimensional (36)
considerations.
The rate of change of density is computed at each
! timestep. From the above definitions, the characteristic
dK 1 m
= m! i Vi 2 − P
& − K e velocity is computed as
dt
& 2"$"
% # Pr oduction %$m#
Change Mean TKE Exhausted 1
MKE MKE MKE   Vp  
2 2
(27)
Vch =  U + u + 
2 ' 2  
dk !e
m   2  
= P
& − m &ε − k + A
&   (37)
dt
& Pr oduction Dissipatio n m
%$# Source
Change TKE TKE Exhausted TKE The heat transfer coefficient can be computed from
TKE TKE
(28) correlations of Nusselt number as a function of Reynolds
number. The interested reader can refer to Assanis and
 3 

ε=
u
 2k
'3
=
2 
3m 
( ) Heywood [15] for details.

'  '  METHOD OF SOLUTION


 
  (29)
The mathematical model results in a set of ordinary
The turbulence production is modeled based on differential equations. These equations must be
2
integrated simultaneously. Due to the incorporation of
2 Cµ k
U chemical kinetics, two different integrators are used to
P = µ t Cβ   ; µ t = (30)
L me ensure computational efficiency. When the chemical
reactions are frozen, the set of equations is integrated
where Cµ=0.09(Universal) and Cβ is an adjustable using the standardized, non-stiff, predictor-corrector
constant. technique, ODERT [21], based on a modified form of the
Large scale eddies can be represented by a geometric Adams Pece method. When pre-combustion reactions
length scale defined as begin to occur late in the compression stroke, the

5
Downloaded from SAE International by University of Michigan, Saturday, July 28, 2018

program switches integrators. A "stiff" ODE integrator, demonstration purposes. Likewise, natural gas, which
DVODE, is then used to deal with the kinetic equations contains a composition on the order of 85-90% methane,
that involve a wide range of reaction time scales, until the 2-10% ethane, and 1-4% propane has received attention
species concentrations are no longer changing. A in recent years due to its "clean burn" potential. Its
flowchart showing the structure of the engine simulation chemical kinetic mechanism is reasonably well
is shown in Figure 2. characterized making it a good candidate for
fundamental studies that can lead to a working engine
After initializing the state variables, the program proceeds
concept. To demonstrate the flexibility of the simulation
through each engine process. The main program
to accept alternative kinetic schemes, the hydrogen
controls the solution process. During each timestep, the
chemistry was described with a reduced scheme
main program calls the pertinent integrator, which then
available in CHEMKIN that considers 11 species and 25
calls one of the four main process routines. Each of
reactions [16]. For natural gas, the GRI-Mechanism 3.0
these process routines determines what equations are
was input into CHEMKIN; the detailed mechanism
important and integrate them over each timestep. The
considers upwards of 53 species and 325 reactions [24].
computer model has been implemented in the FORTRAN
77/90 computer language.
Combustion in a Variable Volume Plenum – Before using
the mechanism in the full simulation, it was studied in a
BEHAVIOR OF SIMULATION SUBMODELS fundamental variable volume, adiabatic reactor to verify
that it reproduced reported ignition trends for variations in
COMBUSTION – The combustion process in the HCCI
inlet temperature, inlet pressure, and the addition of
engine is driven by chemical kinetics. For demonstration
higher order hydrocarbons. Correctly quantifying the
purposes both hydrogen and natural gas were used as
onset of ignition is critical to developing a predictive
the fuels within the full single cycle simulation. Many
simulation since it has a large effect on engine output.
researchers consider hydrogen as the ultimate fuel of the
For our study, the time of occurrence of the peak
future [22]. It has been under study for hybrid applications
concentration of either OH or HO2 was used as the
due to its high energy content and low emissions
ignition criterion. This is consistent with Glassman [25]
potential. In addition, hydrogen exhibits unique fuel
who notes that that the presence of H, O, OH, and HO2
characteristics, which include high flame speed,
abstract H radicals from methane and hence promote
ignitibility of lean mixtures, and high effective octane [23].
chain propagation. Note however, that other studies
Furthermore, its well understood kinetic scheme, which
define the onset of ignition based on the occurrence of
involves only a moderate number of reactions, makes it a
the maximum pressure gradient [25]. Even though
good candidate for modeling studies that can be used for

Start Read Input

Initialize

Integrate Intake 1. Engine Geometry


Main Program 2. Valve Data
Integrate Compression
IVO --- IVC 3. Flow Rates
4. Heat Transfer
Integrate Combustion
IVC --- Ignition 5. Hydro-Carbon Formation
Integrate Expansion 6. Engine Performance
Ignition -- EVO 7. Thermodynamic Properties
Integrate Exhaust 8. Transport Properties
EVO -- IVO
Write Output
Y N
N Y
Combustion
Converge

Finite Rate Specified


Kinetics ROHR
End

Figure 2. Logic Structure of the Thermo-kinetic HCCI Engine Simulation

6
Downloaded from SAE International by University of Michigan, Saturday, July 28, 2018

mixture conditions throughout the compression process 0.022


change until the onset of ignition occurs, a useful
Correlation Coefficient (R) = 1.
quantitative indicator of induction delay (sec) can be

Induction Delay Based on dp/dt (Sec)


0.021
defined as the time that lapses between Bottom Dead
Center (BDC, start of compression) and the onset of
0.02
ignition. Figure 3 examines the correlation between
induction delays based on maximum pressure gradient
0.019
and peak OH or HO2 concentrations for methane
combustion in a variable volume combustor over a range
of inlet temperatures and pressures. The compression 0.018
ratio is 15 and the speed is 1500 rpm. Remarkably, the
correlation coefficient between the two induction delays is 0.017
1, suggesting that both ignition criteria are consistent for
the case of HCCI combustion. 0.016

Having defined the ignition criterion, induction delay


0.015
studies were carried out as the initial conditions of the
0.015 0.016 0.017 0.018 0.019 0.02 0.021 0.022
mixture in the variable volume combustor were varied. Induction Delay based on OH or Ho2 (Sec)
The first case study varied the initial (BDC) temperature
of a methane-air mixture from 400 K to 800 K. The Figure 3. Correlation between induction delays based
equivalence ratio was maintained at 0.5, and the initial on maximum pressure gradient and peak OH
mixture pressure at BDC was 1.5 bar. The resulting or HO2 concentrations for methane
pressure traces are shown in Fig. 4. As anticipated, the combustion.
induction delay decreased with increasing temperature
(see Fig. 5). It was further noted that methane did not
ignite for an intake manifold temperature of 400 K, 1.2 10
8

consistent with previous studies [eg., 10]. 400 K


8 450 K
Next the initial pressure at BDC was varied from 1 bar to 1 10
500 K
3 bar, while keeping the temperature fixed at 500 K, the
600 K
Pressure (Pa)

compression ratio fixed at 15, and the mixture 8 10


7
800 K
composition as above. As shown in Fig. 6, ignition time is
advanced with higher inlet boost. This is a result of the 7
6 10
increased trapped air and fuel concentrations at these
higher inlet pressures generating more free radicals that 7
in turn accelerate the chemical reactions. 4 10

Composition of natural gas can vary widely from state to 7


2 10
state, as well as slightly from cycle to cycle [26]. As a
result many fundamental studies have been carried out to
understand how ignition behaves under varying 0
composition [9, 10, 27, 28]. Compositions were chosen in 0.012 0.014 0.016 0.018 0.02 0.022 0.024 0.026
a range consistent with Westbrook and Pitz [27]. A Time (Sec)
methane-air mixture at 1.5 bar and a temperature of 500 Figure 4. Effect of initial mixture temperature on ignition
K was used as the baseline case. Ethane and then and pressure profile in a variable volume
finally propane was gradually added to the mixture in chamber. Methane combustion with
increased quantities. Computed induction delays as a equivalence ratio of 0.5 and initial mixture
function of composition are shown in Fig. 7. Consistent pressure of 1.5 bar.
with Westbrook and Pitz [27], Fraser et. al. [28], Naber et.
al. [9], and Agarwal and Assanis [10], increased
concentrations of ethane and propane reduced induction
delays, which is a result of increased radical pools
promoting chain branching reactions.

7
Downloaded from SAE International by University of Michigan, Saturday, July 28, 2018

0.022 the engine geometric compression ratio was fixed at 15,


(iii) intake manifold pressure was 1.5 bar, intake manifold
temperature was 425 K, mixture equivalence ratio was
0.021
0.3. and (iv) engine speed was 1500 rpm.
0.02 0.0193
Induction Delay (Sec)

Baseline: 1.0 CH 4
0.019
0.0192

0.018

Induction Delay (Sec)


0.0191
0.017 0.97 CH 4 + 0.03 C2H6
0.019
0.016
0.95 CH4 + 0.05 C2H6

0.0189
0.015
0.0012 0.0014 0.0016 0.0018 0.002 0.0022 0.0024 0.90 CH 4 + 0.10 C2H6
-1
(Temperature) 1/K 0.0188

Figure 5. Correlation of induction delay as a function of 0.80 CH4 + 0.10 C2H6 +0.1 C3H8
initial mixture temperature. Methane 0.0187
combustion in a variable volume chamber 0 1 2 3 4 5 6
with equivalence ratio of 0.5 and initial Test Case
mixture pressure of 1.5 bar.
Figure 7. Variation in induction delay under different
compositions of methane, ethane and
8
2.5 10 propane. Natural gas combustion in a
1 BAR variable volume chamber with initial mixture
conditions of 1.5 bar and 500 K, and an
8
2 10 1.5 BAR equivalence ratio of 0.5.
2 BAR
Pressure (Pa)

8 Table 1 Engine Specifications


1.5 10 3 BAR
Engine Type 4-Stroke
Displacement 12.7 Liter
8 Bore 13cm
1 10
Stroke 16cm
Con. Rod Length 26.93cm
7 Intake Valve Opening 685 deg *
5 10
Intake Valve Closing 220 deg *
Exhaust Valve Opening 485 deg *
Exhaust Valve Closing 795 deg *
0
* Relative to TDC Intake
0.01 0.015 0.02 0.025 0.03
Time (Sec)
A pressure-volume diagram, over the full engine cycle, is
Figure 6. Variation of onset of ignition with changing shown in Fig. 8a. Clearly, the combustion is very close to
initial pressure of mixture. Methane constant volume. This rapid rate of heat release is in
combustion in a variable volume chamber agreement with the work of Van Blarigan and
with equivalence ratio of 0.5 and initial Goldsborough [5]. The pressure and temperature history
mixture temperature of 500 K. is shown in Fig. 8b. As will be shown in subsequent
sections, the rapid variation of ignition as well as rapid
Combustion in Engine Simulation – The behavior of the heat release greatly effect engine performance. The
combustion submodel was further explored within a full species formed during hydrogen combustion are tracked
engine cycle simulation. Primary engine specifications as a function of time, and are shown in Fig. 9. Peaks in
are summarized in Table 1. Unless otherwise noted, for H, OH, and HO2 can be discerned near the ignition point.
the engine combustion study reported in this section, as Note that NO emissions are near zero, a result of lean
well as for subsequent studies illustrating the behavior of combustion.
other submodels: (i) hydrogen was used as the fuel, (ii)

8
Downloaded from SAE International by University of Michigan, Saturday, July 28, 2018

1000 0.0015 0.0006


OH H2O2
(a)
0.0005

Major Species Mole Fractions (-)


HO2

Minor Species Mole Fractions (-)


N
100 0.001
O 0.0004
NO
LN Pressure (bar)

HO2, H, and OH, production


0.0003
precursor for ignition
10 0.0005
0.0002

0.0001
1 0

0
NO Production
-0.0005 -0.0001
0.1

0.0085

0.0095

0.0105

0.0115
0.008

0.009

0.011

0.012
0.01
0.0001 0.001 0.01
3
LN Volume (m ) Time (Sec)

120 2500 Figure 9. Variation of major and minor species during


hydrogen combustion in an engine operating
(b) with an inlet manifold temperature of 425 K,
100
2000 an inlet pressure of 1.5 bar, and an
equivalence ratio of 0.3.
80
Temperature (K)
Pressure (bar)

Ignition Point 1500 The first is due to the rapid blowdown event that occurs
when the exhaust valve opens, and the second is
60
Temperature consistent with the piston motion. Figure 10b shows the
1000 calculated mean flux velocities for both the intake and
40
Pressure
exhaust velocities. The large flow velocities have a direct
effect on the in-cylinder flow field and resulting heat
500 transfer coefficient. Note that negative velocities
20
correspond to reverse flows due to an adverse pressure
gradient across the valve.
0 0
200 250 300 350 400 450 500 Including the gas exchange processes within the cycle
Crank Angle (deg) simulation, as opposed to specifying thermodynamic
conditions at BDC, allows the simulation to be used for
Figure 8. Cyclic characteristics of a four-cycle engine variable valve timing studies. For demonstration
running on hydrogen at an inlet manifold purposes, the exhaust valve closing time was varied from
temperature of 425K, an inlet pressure of 1.5 725° to 795° ca-deg to demonstrate the effect that it has
bar, and an equivalence ratio of 0.3. (a) on ignition and the resulting engine performance. The
pressure-volume diagram; (b) variation of gas test case was run using hydrogen fuel, an inlet pressure
pressure and temperature during combustion. of 1.5 bar, and an inlet manifold temperature of 500K,
although the results are qualitatively the same at other
GAS EXCHANGE – Profiles detailing the rate of mass operating conditions. Figure 11 shows results for ignition
flow through the engine intake/exhaust valves, computed timing, indicated thermal efficiency, and volumetric
using the one-dimensional, quasi-steady, compressible efficiency. Closing the exhaust valves early results in a
flow model, are shown in Fig. 10a. It is noted that the reduction in scavenging efficiency. This causes more
intake valve mass flow rate follows the path of the piston internal EGR to reside in the cylinder after the exhaust
motion, in agreement with Assanis and Heywood [15]. On valves close, and leads to higher mixture temperatures
the other hand, the exhaust flow exhibits two peaks after induction. The elevated temperatures and active
radicals in the internal residual gas cause earlier ignition
timing. Since ignition timing is advanced into the
compression stroke, the thermal efficiency is greatly
reduced with earlier ignition.

9
Downloaded from SAE International by University of Michigan, Saturday, July 28, 2018

1.2 TURBULENT FLOW MODEL – The eddy length scale of


(a)
the turbulence is governed as a function of the distance
maximum 1 between cylinder head and piston, as noted in Eq. (33).
Intake Exhaust Its variation through the engine cycle is shown in Figure
0.8 12a. More details can be found in Assanis and Heywood
Mass Flow/Mass Flow

[15]. Figure 12b shows the variation in the mean flow


0.6 Exhaust Blowdown velocity, turbulent intensity, and the overall characteristic
velocity. It is noted during the intake stroke that the mean
0.4 kinetic energy is initially large and then decreases in
response to piston motion. It decreases more slowly
0.2 during the compression stroke. The mean flow then
increases during the exhaust blowdown, a result of the
0
large pressure potential across the valve. The large
scale turbulence levels are shown to be very high during
-0.2
-100 0 100 200 300 400 500 600 700
the intake flow. Later in the intake event, turbulence
Crank Angle (deg) decays since the rate of dissipation, which is governed by
a scaling argument, due to viscous shear stresses is
400
larger than the rate of turbulence production. The use of
Exhaust Valve Velocity
(b) RDT to account for rapid density changes results in a
300 slight amplification of the turbulence intensity during
combustion. The characteristic velocity is a scaling
Intake Valve Velocity
Valve Velocity (m/s)

200 argument made up of mean piston speed, mean gas


velocity, turbulent intensity, and during the exhaust, a
100
blowdown velocity. The variation of both velocity flow
scales as well as the piston speed, has a direct effect on
the characteristic velocity as well as the subsequent heat
0
transfer process.

-100 HEAT TRANSFER MODEL – Once the characteristic


velocity is found, the film coefficient for heat loss can be
Reverse Flow To Intake Manifold
-200 calculated. The heat transfer coefficient and heat transfer
-100 0 100 200 300 400 500 600 700 per unit area are shown in Figure 13. It is noted that it
Crank Angle (deg) experiences a strong variation over the cycle. This is due
Figure 10. (a) Mass flows and (b) velocities through the to a combination of variation in gas properties (i.e.
intake and exhaust valves for hydrogen-fueled thermal conductivity and viscosity) as well as the
engine operating at 1500 rpm. characteristic velocity and the macroscale of turbulence.
The second peak in the film coefficient is due to the large
120 360 exhaust blowdown.
Ind.Thermal Efficiency, Volumetric Efficiency (%)

Volumetric Efficiency A heat transfer study was performed to examine the


100
effect of different heat transfer models on ignition.
355
Specifically, the Woschni heat transfer model [13], that
has been used in the majority of published HCCI studies,
Ignition Point (deg)

80 350 was compared to the zero-dimensional k-ε based heat


transfer model by Assanis and Heywood [15]. The
adiabatic combustion case was used as a baseline in the
60 Ignition Point 345 study. The comparisons were run at a compression ratio
of 17, an inlet pressure of 3 bar, and an inlet temperature
of 450 K, although similar trends follow at other operating
40 340 conditions. The pressure and temperature profiles are
Thermal Efficiency shown in Figs. 14a and 14b. Clearly, the heat transfer
model has a pronounced effect on the ignition point. As
20 335
720 730 740 750 760 770 780 790 800
anticipated, the adiabatic case ignites first because the
Exhaust Valve Closing (deg)
lack of heat transfer promotes a higher mean cylinder gas
temperatures and hence earlier ignition. But what is
Figure 11. Variation of volumetric efficiency, thermal most surprising is that the Woschni model predicts the
efficiency, and ignition timing for different ignition point 8 degrees later than that by Assanis and
exhaust valve closing times. Hydrogen-fuelled Heywood [15]. This discrepancy is attributed to the fact
engine operating with intake manifold that the Woschni model was developed for DICI diesel
pressure of 1.5 bar and intake manifold engines by fitting data to the correlation form. In other
temperature of 500 K.
10
Downloaded from SAE International by University of Michigan, Saturday, July 28, 2018

words, it is a linearized convective/radiative model 20 00 10 0


specifically applicable to non-premixed combustion.
0.6 H eat Transfer C o effici ent R ate of H ea t T ransfer

H eat Transfer C o effici ent (w / m K)


80

R ate of H e at T ran sfe r (k W )


(a) 15 00
Turbulence Length Scale/Piston Bore

0.5
60

0.4 10 00

40
0.3 E xhaust B lo wd ow n
50 0
20
0.2

0.1 0 0
TDC 10 0 20 0 30 0 40 0 50 0 60 0
C rank A ngle (d eg)
0
-100 0 100 200 300 400 500 600 700 Figure 13. Cyclic variation of heat transfer coefficient
Crank Angle (deg) and gas-to-wall heat transfer rate for a
hydrogen-fuelled engine at 1500 rpm, an inlet
pressure of 1.5 bar, an inlet temperature of
140
(b) 425 K and an equivalence ratio of 0.3.
120
PARAMETRIC STUDIES WITH ENGINE
Flow Velocity (m/sec)

100 Mean Flow Velocity SIMULATION


Turbulent Intensity
Characteristic Velocity Having demonstrated the behavior of the physical
80
submodels in the four-stroke HCCI engine cycle
Exhaust Blowdown simulation, several parametric studies were performed to
60
Turbulence Amplification
explore the effect of intake mixture conditions, as well as
40 geometric and effective compression ratios on engine
performance and efficiency. These parametric studies
20 have been carried-out for the engine specifications listed
in Table 1, but for geometric compression ratios as
0 specified. Both hydrogen and methane have been used
-100 0 100 200 300 400 500 600 700 as the fuels for these engine studies. The operating
Crank Angle (deg) speed was 1500 rpm. Integrated cycle results are
interpreted in the light of cycle-resolved pressure and
Figure 12. Cyclic variation of (a) integral length scale;
temperature profiles, as well as onset of ignition.
(b) mean flow velocity, turbulent intensity, and
characteristic velocity at 1500 rpm.
TEMPERATURE SWEEP – The first study was run with
Hydrogen-fuelled engine with an inlet
hydrogen at a compression ratio of 15:1, an inlet manifold
pressure of 1.5 bar, an inlet temperature of
pressure of 1.2 bar, and an equivalence ratio of 0.3. The
425 K and an equivalence ration of 0.3.
inlet manifold temperature was varied between 400 K and
800 K, and the effect on the ignition point was studied. As
Figure 14c shows the heat transfer coefficients for the
shown in Fig. 15a, when temperature is increased, it
Woschni correlation and that predicted by the k-ε model.
advances the onset of ignition due to increased reaction
The Woschni coefficient shows very little variation over
rates. However, it is also important to note how
the gas exchange event, when heat transfer should be
increasing inlet manifold temperatures negatively affect
higher due to mean gas flow and the production of large-
indicated thermal efficiency, volumetric efficiency, and
scale turbulence. By including velocity scales for both the
trapped mass. The trends captured in Fig. 15b reflect
mean flow and turbulence intensity, Assanis and
that an increase in the inlet temperature reduces trapped
Heywood [15] predict an elevated film coefficient during
mass and volumetric efficiency, which in turn adversely
the gas exchange processes. Over the combustion event
affects torque and power output. Advanced ignition,
the Woschni model predicts a film coefficient on the order
which increases compression effort, combined with
of three times larger than the peak value of the k-ε model.
reduced volumetric efficiencies leads to the observed
This is again attributed to radiation being indirectly
reduction in net indicated thermal efficiency. It is also
present in the Woschni predictions.
noted that one of the cases did not

11
Downloaded from SAE International by University of Michigan, Saturday, July 28, 2018

300 12 0

(a) 400 K (a )
250 10 0
425 K

450 K
200 80 500 K
Pressure (bar)

P res su re (b a r)
600 K

150 60

100 40

Adiabatic, IGN = 354.5


50 k-ε model, IGN = 368.1 20

Woschni, IGN = 376.1


0 0
340 350 360 370 380 390 400 300 320 340 36 0 3 80 40 0
Crank Angle (deg) C ra nk An g le (d e g)

2500 120 2.6

(b) (b)
100 2.5
2000
Volumetric Efficiency 2.4

Thermal Efficiency (%)


80

Mass Trapped (gram)


Mass Trapped
Temperature (K)

1500 2.3
60
Thermal Efficiency
2.2
1000 40
2.1
No Ignition
20
Adiabatic, IGN = 354.5 2
500
k- ε model, IGN = 368.1
0 1.9
Woschni, IGN = 376.1

0 -20 1.8
340 350 360 370 380 390 400 350 400 450 500 550 600 650
Crank Angle (deg) Temperature (K)
4
2 10 Figure 15. Effect of inlet manifold temperature on (a)
k- ε m o d el
ignition timing and cyclic pressure profile, and
(c)
W o sch n i m o d el (b) Thermal efficiency, mass trapped and
H e a t Tra n sfe r C o e ffici e nt (W /m K)

4 volumetric efficiency. Hydrogen-fuelled


2

1 .5 1 0
engine with intake manifold pressure of 1.5
bar and equivalence ratio of 0.3.
4
1 10 ignite, therefore demonstrating how the simulation can be
used to map an engine's flammability limit.

5 00 0
EQUIVALENCE RATIO SWEEP – The next study was
conducted with hydrogen fuel for a range of equivalence
ratios between 0.15 and 0.4, at an inlet manifold pressure
of 1.2 bar and a temperature of 400 K. The engine
0 compression ratio was maintained at 15. What was
0 10 0 2 00 3 00 4 00 500 600
noted is that increased fuel concentrations promoted
C ra n k An g le (d e g)
radical generation and thus led to shorter induction
Figure 14. (a) Pressure, (b) temperature, and (c) heat delays. The increased heat release associated with
transfer coefficient profiles predicted by k-ε elevated equivalence ratios promotes higher cylinder
heat transfer model, Woschni heat transfer pressures and specific power outputs, up to the point that
model, and assuming adiabatic operation. its benefits are offset by the increased compression work
HCCI engine with CR=17, operating on caused by overadvanced ignition. The simulation
natural gas, with an equivalence ratio of 0.4, demonstrates how the lean flammability limit of the
intake manifold pressure of 3.0 bar, and engine can be investigated. In this case, note from Fig.
intake manifold temperature of 450 K.

12
Downloaded from SAE International by University of Michigan, Saturday, July 28, 2018

16 that hydrogen mixture fails to ignite at an equivalence induction delay increases as geometric compression ratio
ratio of 0.15. is decreased. Figure 18, which shows temperature
profiles for the three cases at 420 K, clearly shows that
10 0
temperature rises at a slower rate in the case with the
φ = 0 .1 5
lowest mechanical compression ratio. Since the
80 φ = 0 .2
Arrhenius kinetic rates exhibit an exponential
φ = 0 .3 temperature dependence, a lower temperature profile will
φ = 0 .3 5 contribute to a longer induction delay as a result of slower
Pre ssu re (b ar)

60 φ = 0 .4 rates of radical generation.

370 14
40

Indicated Mean Effective Pressure (bar)


ID (Tman = 420 K)
12
365
ID Tman = 450 K
20 Pmanifold = 3.0 bar 10

Induction Delay (deg)


360

8
0
355
30 0 32 0 340 36 0 380 400
6
C ra nk An g le (d e g)
350 P manifold = 2.0 bar
Figure 16. Effect of equivalence ratio on ignition timing P manifold = 1.49 bar 4
and cyclic pressure profile. Hydrogen-fuelled IMEP (420 K)
engine with intake manifold pressure of 1.2 345
2
IMEP (450 K)
bar and intake manifold temperature of 410 K.
340 0
COMPRESSION RATIO SWEEP – One of the criticisms 12 16 20 24 28 32 36 40 44
of the lean burn HCCI concept is that it lends itself to a Mechanical Compression Ratio
very efficient, but low power output operation. This is in
Figure 17. Effect of mechanical compression ratio on
part due to the fact that extremely high inlet manifold
induction delay and mean effective pressure.
temperatures are often needed to promote ignition of
Methane-fuelled engine, operated at three
natural gas at top dead center (TDC). As a result the
intake manifold pressures (3.0, 2.0 and 1.49
volumetric efficiency and overall engine power output are
bar) and two intake manifold temperatures
adversely affected. To circumvent this problem, a high
(420 K and 450 K), all at an equivalence ratio
compression ratio can be used to elevate cylinder
of 0.4.
pressures and temperatures. As a result, the fuel-air
mixture would not have to be heated to extraordinarily
high levels to promote ignition near TDC, and the 2500
elevated expansion ratios would promote increased
thermal efficiencies.
2000
Realizing that both geometric compression ratio and inlet
boost level contribute to the effective compression ratio
Temperature (K)

that the mixture is exposed to, it is important to 1500


investigate the optimum levels for each. To address this
issue, a simulation case study was performed for different 20:1 Compression Ratio
1000
geometric compression ratios, but modulated inlet boost
pressure levels to yield a constant effective compression
ratio. The latter was defined to produce the same peak
500 35 : 1 Compression Ratio
motoring pressure under all conditions. The chosen 28 : 1 Compression Ratio
geometric compression ratio set included 20, 28, and 20 : 1 Compression Ratio
35:1. The corresponding inlet pressures were adjusted 0
to values of 3, 2.5, and 1.49 bar, respectively. This 320 330 340 350 360 370 380 390 400
exercise was carried out for two levels of intake manifold Crank Angle (deg)
temperature, i.e. 420 K and 450 K.
Figure 18. Cyclic gas temperature profiles for three
As shown in Fig. 17, the lowest geometric compression different mechanical compression ratios, but
ratio resulted in the highest Indicated Mean Effective the same effective compression ratio..
Pressure. This is a result of the increased trapped mass Methane-fuelled engine at an intake manifold
caused by the higher boost level used in conjunction with temperature of 420K and an equivalence ratio
the lowest geometric compression ratio. Notice also that of 0.4.

13
Downloaded from SAE International by University of Michigan, Saturday, July 28, 2018

Since the case with an inlet pressure of 3 bar exhibited relative to TDC, causes elevated pressures during the
the highest efficiency, a parametric study was performed piston upstroke. This increases the compression work,
to optimize geometric compression ratio at this level of which in turn reduces the net piston work, as well as
inlet boost. To promote ignition at lower compression increases peak cylinder pressures dramatically, from 190
ratios, inlet temperature was elevated to 500 K. The bar to approximately 330 bar. Consequently, IMEP is
equivalence ratio was held constant at a lean value of reduced from 11.5 bar to under 10 bar, and brake thermal
0.4 and the compression ratio was swept from 12 to 24:1. efficiency drops from over 50% to 44% (see Fig. 19d).
The resulting pressure profiles, near TDC, are shown in Figure 19c also depicts the behavior of IMEP non-
Fig. 19a. First, it is noted that the increased geometric dimensionalized with peak cylinder pressure as a
compression ratio promotes a reduction in the induction function of the timing of ignition. This normalized ratio,
delay, as shown in Fig. 19b. This is a result of more essentially a measure of power output per unit
elevated temperature histories earlier in the cycle, which mechanical stress, is reduced by a factor of 2 as
promote chain-initiating reactions earlier. The timing of geometric compression ratio is increased from 12 to 24.
occurrence of ignition greatly affects engine power output This indicates a disproportionate stress penalty for an
as well as mechanical stresses. Figure 19c correlates HCCI engine where the onset of ignition occurs
the timing of ignition with Indicated Mean Effective prematurely. This case study suggests that a geometric
Pressure (IMEP). Early ignition of the fuel-air charge, compression ratio of 17:1 with an inlet boost of 3 bar can

350 11.5 7
CR 12 :1
IMEP (c)
300 CR 15:1 (a) 11
CR 17:1 6
CR 20:1
250 10.5
CR 22:1

100*IMEP/P
IMEP (bar)
Pressure (bar)

CR 24:1 5
200 10

IMEP/P
150 9.5 max

max
4

(bar)
100 9
3
50 8.5

0 8 2
320 340 360 380 400 345 350 355 360 365 370
Crank Angle (deg) Ignition Timing (deg)

182 11.5 51

180 Thermal Efficiency


(b) (d) 50
11
178
Brake Thermal Efficiency
49
Induction Delay (Sec)

176
10.5 48
BMEP
174
BMEP (bar)

47
172 10
46
170
45
9.5
168
44
166
9 43
15 20 25 14 16 18 20 22 24 26
Compression Ratio Compression Ratio

Figure 19. Effect of mechanical compression ratio on (a) cyclic pressure profile, (b) induction delay, (c) ignition timing and
IMEP, and (d) BMEP and brake thermal efficiency. Methane-fuelled engine at an intake manifold temperature of 500K, an
intake manifold pressure of 3.0 bar and an equivalence ratio of 0.5.fig 19

14
Downloaded from SAE International by University of Michigan, Saturday, July 28, 2018

produce reasonable power outputs in an HCCI 4. Several parametric studies of inlet temperature,
configuration, while limiting peak cylinder pressures. pressure, and equivalence ratio were carried out
This finding is in line with tests that have been performed using the complete cycle simulation. It was noted
by Christensen et. al., [29]; and simulation work [7]. More how mixture preparation as well as the ensuing
importantly though, this case study has demonstrated ignition point greatly affect the volumetric efficiency
how the four-stroke thermokinetic simulation can be used and the power output. It was also concluded that
to define efficient operating ranges for the HCCI engine. HCCI operation with Natural Gas needs high
compression ratios to promote ignition with reduced
CONCLUSIONS intake temperatures.
5. A case study was performed for varying geometric
A computer simulation of the Homogenous Charge compression ratios, but at constant effective
Compression Ignition (HCCI) engine has been developed compression ratios achieved by modulating inlet
for ignition and performance studies. The simulation manifold pressure. It was noted that the lowest
couples models for mass, species, and energy within a mechanical compression ratio and highest inlet boost
zero-dimensional framework. A methodology has also produced the highest efficiency and power output.
been developed for integrating CHEMKIN libraries with This is a result of ignition timing occurring after TDC.
physical models of engine processes in the context of a due to a lag in the gas temperature.
full cycle simulation. Parametric studies have been
6. A follow-up study was then performed using an inlet
conducted to illustrate the behavior of the combustion
pressure of 3 bar over varying geometric
model, as well as the other physical process models.
compression ratios. The tradeoffs associated with
The cycle simulation has been exercised to study ignition
power and mechanical stress suggest that a 17:1
and performance of natural-gas and hydrogen fuelled
compression ratio with 3 bar of boost lends itself to
HCCI engines as a function of operating and design
reasonable power outputs for the HCCI, Natural Gas
conditions. The following conclusions have been drawn
concept.
from our study:
Overall, this study demonstrates the importance of
1. It has been demonstrated that the widely available
coupling detailed chemistry descriptions with physical
and used CHEMKIN libraries can be effectively
models of the rest of the HCCI engine processes, notably
integrated within a full cycle simulation of the HCCI
gas exchange and turbulent heat transfer.
engine. This enables the user to select from different
fuels, and readily describe their chemical kinetics
ACKNOWLEDGEMENTS
using reported simplified or complex chemical
schemes, as appropriate. Use of CHEMKIN also
The authors gratefully acknowledge a gift from
makes it possible to take advantage of its integrator
Caterpillar, Inc. to the University of Michigan that was
DVODE to solve stiff chemical reaction systems.
used to partially support this study.
2. For illustration, two chemical kinetics schemes
describing hydrogen and natural gas chemistry have REFERENCES
been implemented in the code. The hydrogen
scheme is a reduced one, consisting of 11 species 1. Onishi, S., S. J. Hong, K. Shoda, P. Do Jo, and S.
and 23 reactions. The natural gas chemistry is Kato, 1979, “Active Thermo Atmospheric
described via the GRI-Mech 3.0 that considers 53 Combustion (ATAC) – A New Combustion Process for
species and 325 reactions, including NOx chemistry. Internal Combustion Engines,” SAE Paper 790501,
The behavior of the complex chemistry scheme was 1979.
first studied in a variable volume bomb. Predicted 2. Najt, P.M. and D.E. Foster, “Compression-Ignited
ignition trends were consistent with those reported in Homogeneous Charge Combustion,” SAE Paper
the literature as temperature, pressure, equivalence 830264, 1983.
ratio, and composition were varied. 3. Thring, R.H., “Homogeneous Charge Compression
3. It has been demonstrated that predicting heat Ignition (HCCI) Engines,” SAE Paper 892068, 1989.
transfer losses based on velocity and length scales 4. Van Blarigan, P., N. Paradiso, and S. Goldsborough,
that are physically connected to in-cylinder “Homogeneous Charge Compression Ignition with a
processes (gas exchange, mean flow, turbulence, Free Piston: A New Approach to Ideal Otto Cycle
piston motion) are substantially different than those Performance,” SAE Paper 982484, 1998.
based on the empirical, Woschni heat transfer 5. Van Blarigan, P., and S. Goldsborough, “A Numerical
correlation. It is felt that the experimental database Study of a Free Piston Engine Operating on
on which the latter is based is not appropriate for Homogeneous Charge Compression Ignition
HCCI engines. It is also shown that differences in Combustion,” SAE Paper 990619, 1999.
heat transfer predictions can shift the predicted start 6. Ryan, T.W., and T. Callahan, “Homogeneous Charge
of ignition by up to 10 deg. Compression Ignition of Diesel Fuel,” SAE Paper
961160, 1996.

15
Downloaded from SAE International by University of Michigan, Saturday, July 28, 2018

7. Smith, J.R., S.M. Aceves, C. Westbrook, and W. Pitz, 21. Shampine, L.F., and M.K. Gordon, Computer
“Modeling of Homogeneous Charge Compression Solution of Ordinary Differential Equations: The Initial
Ignition (HCCI) of Methane,” Proceedings of the Value Problem, Freeman, 1974.
1997 ASME Internal Combustion Engine Fall 22. Maxwell T.T. and J.C. Jones, Alternative Fuels:
Technical Conference, ASME Paper No. 97-ICE-68, Emissions, Economics, and Performance, Society of
ICE-VOL. 29-3, pp. 85-90, 1997. Automotive Engineers, Inc., Warrendale, PA, 1990.
8. Gray, A.W., and T.W. Ryan, “Homogeneous Charge 23. Blarigan P.V., “Development of a Hydrogen Fueled
Compression Ignition of Diesel Fuel,” SAE Paper Internal Combustion Engine Designed for Single
971676, 1997. Speed/Power Operation”, SAE Paper 961690, 1996.
9. Naber J.D., D.L. Siebers, J.A. Caton, Westbrook C.K., 24. Smith, G.P., G.M. Golden, M. Frenklach, N.W.
Di Julio S.S., “Natural Gas Autoignition Under Diesel Moriarty, B. Eiteneer, M. Goldenberg, T. Bowman, R.
Conditions: Experiments and Chemical Kinetic Hanson, S. Song, G.C. Gardiner Jr., V. Lissianski,
Modeling,” SAE Paper 942034, 1994. and Z. Qin, GRI-MECH 3.0, http://
10. Agarwal, A. and D.N. Assanis, “Modeling the Effect www.me.berkeley.edu/gri_mech/
of Natural Gas Composition on Ignition Delay Under 25. Glassman, I., Combustion, Academic Press, Inc.,
Compression Ignition Conditions,” SAE Paper San Diego, Ca, 1996.
971711, 1997.
26. Liss, W.E., and W.H. Thrasher 1991, “Natural Gas as
11. Aceves, S.M., Smith, J.R., Westbrook, C.K., Pitz, a Stationary Engine And Vehicle Fuel,” SAE Paper
W.J., “Compression Ratio Effect on Methane HCCI 912364.
Combustion,” Journal of Engineering for Gas
Turbines and Power, Vol. 121, pp. 569-574. 27. Westbrook C.K., and Pitz W., 1983, “Effects of
Propane and Ignition of Methane-Ethane-Air
12. Kusaka, J., T. Yamamoto, Y. Daisho, R. Kihara, T. Mixtures,” Combustion Science and Technology, Vol.
Saito, and O. Shinjuku, “Predicting Homogeneous 33, pp. 315-319.
Charge Compression Ignition Characteristics of
Various Hydrocarbons,“ Proceedings of the 15th 28. Fraser, A.F., Siebers, D.L., Edwards, C.F.,
Internal Combustion Engine Symposium “Autoignition of Methane and Natural Gas in a
(International), Seoul, Korea, 1999. Simulated Diesel Environment,” SAE Paper 910227,
1991.
13. Woschni, G., “A Universally Applicable Equation for
the Instantaneous Heat Transfer Coefficient in the 29. Christensen, M. and Johansson B., “Influence of
Internal Combustion Engine,” SAE Paper 670931, Mixture Quality on Homogeneous Charge
1967. Compression Ignition,” SAE Paper 982454, 1998.
14. Poulos, S.G., and Heywood J.B., “The Effect of
NOMENCLATURE
Chamber Geometry on Spark-Ignition Engine
Combustion”, SAE Paper 830334, SAE Trans., Vol.
92, 1983. Symbol Definition
15. Assanis, D. N., and Heywood, J. B., “Development Thermodynamic Properties
and Use of Computer Simulation of the
Turbocompounded Diesel System for Engine P Cylinder Gas Pressure
Performance and Component Heat Transfer Studies,” T Temperature
SAE Paper 860329, 1986. u Internal Energy
16. Kee, R.J., F.M. Rupley, and J.A. Miller, “Chemkin-II: A h Enthalpy
FORTRAN Chemical Kinetics Package for the SH Specific Heat
Analysis of Gas-Phase Chemical Kinetics,” Sandia Cv SH at Const. Volume
National Labs Report SAND89-8009B, 1991. Cp SH at Const. Pressure
17. Van Wylen, G.J., R. Sonntag, R., and C. Borgnakke, R Gas Constant
Fundamentals of Classical Thermodynamics, 4th ρ Density
edition, John Wiley & Sons, 1994.
v Specific Volume
18. Assanis, D.N., and Polishak, M., “Valve Event φ Equivalence Ratio
Optimization in a Spark-Ignition Engine”, ASME
Trans., Journal of Eng. for Gas Turbines and Power, Yj Mass Frac. Species ‘j’
Vol. 112, pp. 341-347, 1990. Wmw Molecular Weight
19. Jessee, J.P., R.F. Gansman, and W.F. Fiveland, Energy Transfers
“Multi-Dimensional Analysis of Turbulent Natural Gas
Flames Using Detailed Chemical Kinetics,” W Displacement Work
Combustion Science and Technology, Vol.129, pp. Qht Heat Transfer
113-140, 1997.
Engine Geometry
20. Agarwal, A., Z. Filipi, D. N. Assanis, and D. Baker,
“Assessment of Single-and Two-Zone Turbulence Ab Piston Bore Area
Formulations for Quasi-Dimensional Modeling of Vd Displacement Volume
Spark-Ignition Engine Combustion, Combustion
Science and Technology, Vol.136, pp. 13-39, 1998.

16
Downloaded from SAE International by University of Michigan, Saturday, July 28, 2018

Chemical Kinetic Parameters


Ea Activation Energy
kf,r F/R Reaction Rates
w!j Mass Rate of Prod.
Ωi Molar Rate of Prod
νi,ii Stoichiometric Coeff.
qi Progress Variable

Heat Transfer Parameters


U Mean Cylinder Velocity
u’ Turbulent Velocity
KE Kinetic Energy
K Mean KE
k Turbulent KE
P Turbulent Energy Prod.
µt Turbulent Viscosity
' Integral Length Scale
A Amplification (RDT)
ε Eddy Dissipation
Vch Characteristic Velocity
kg Thermal Conductivity
hg Film Coefficient

17

You might also like