Sie sind auf Seite 1von 15

International Journal of Mechanical Engineering Education Vol 30 No 4

Illustration of the use of an


instructional version of a
thermodynamic cycle
simulation for a commercial
automotive spark-ignition
engine
JERALD A. CATON, Texas A&M University, Department of
Mechanical Engineering College Station, TX 77840, USA.
jcaton@mengr.tamu.edu
Received 6th December 2000
Revised 30th January 2001
The development and use of an instructional version of a thermodynamic engine cycle simula-
tion for classroom use is described. This simulation is based on well-established features, but
which are not necessarily the most advanced. The major simplification of this instructional
simulation is the use of constant specific heat capacities as opposed to the use of variable
composition and properties. The cycle simulation was developed with an elementary set of
conventional sub-model components. To account for the unsteady flow dynamics, an empirical
adjustment factor was used. With the exception of this empirical adjustment factor, all of the
constants associated with the sub-models are used as suggested by the original publications.
Students, therefore, are readily able to develop and use this simulation.
This paper then demonstrates the usefulness of such a basic simulation in describing the
overall performance of a commercial automotive spark-ignition engine for a range of engine
speeds and operating conditions. A modern, four-valve per cylinder, two-camshaft engine was
selected for this study. Although the cycle simulation was based on elementary conventional
features, a number of important engine characteristics were correctly obtained. These included
the overall performance for engine speeds up to 7000 rpm, and details such as the time (crank
angle) of peak pressure for optimum performance.
Key words: engine cycle simulations, engine thermodynamic models
INTRODUCTION
Over the past forty or so years, engine cycle simulations have been developed and used to
study a variety of features and issues of automotive engines (e.g. [16]). Todays engine cycle
International Journal of Mechanical Engineering Education Vol 30 No 4
284 Jerald A. Caton
simulations are sophisticated, complex computer programs that often require a high-end work-
station or main-frame computer. Many of these simulations contain advanced and detailed sub-
models for the fluid mechanics, heat transfer, friction, combustion and chemical kinetics.
For classroom use, the development of these complete and sophisticated simulations by
undergraduate students in a one-semester course may not be possible. As an alternative to the
development of these rigorous simulations, student development of an instructional version of
these thermodynamic engine simulations may provide an educational role. The goal of these
instructional simulations is not to be completely predictive, but rather to assist in understand-
ing engine operation and in categorizing performance trends. In addition, these instructional
simulations are designed to introduce students to the subject of engine cycle simulations.
Although relatively basic, such simulations still retain most of the important features of
engines. In particular, these features include time (or crank-angle) varying quantities such as
the cylinder gas pressure and temperature, heat release, heat loss, intake and exhaust flow rates,
and cylinder mass. In addition, the basic simulation described here includes parameters which
vary as functions of engine speed such as: combustion duration, friction, exhaust and intake
manifold pressure, and an empirical adjustment factor to account for the unsteady flow
dynamics.
As described elsewhere [7, 8], engine performance results from an instructional version of
the engine cycle simulation are in good agreement with results from a complete version of the
simulation. As might be expected, quantities such as the magnitudes of the peak pressure and
temperature were not well duplicated since the effect of gas properties was missing from the
instructional version. On the other hand, the time (crank angle) of the peak pressure and
temperature were closely matched. As will be demonstrated in this paper, many of the more
important features of engine operation and performance are provided by these instructional
versions of the simulations.
Although the use of basic thermodynamic engine cycle simulations may provide several
advantages due to the reduced computational times relative to more complete simulations, that
is not the motivation for using these instructional versions. These reduced computational times
are a result of assuming constant specific heat capacities (which results in the requirement that
the composition is constant), and to a lesser degree, due to the use of less advanced sub-
models.
The disadvantages of using such a basic cycle simulation is that the lack of time-varying
specific heat capacities results in a loss of accuracy. As mentioned above, the maximum cylin-
der pressure and temperature would be expected to be over predicted without the correct high
temperature instantaneous properties. Additionally, the effect of varying the inlet fuelair ratio
cannot be described in an exact fashion. Finally, the lack of time-varying specific heat
capacities (and the related lack of variable composition) precludes the ability to accurately
predict exhaust emissions.
This paper will first describe the model development, and then illustrate the use of the basic
model with an application to a modern automotive engine. Complete details of the derivation
and descriptions of all the results are provided elsewhere [710].
MODEL DESCRIPTION
This section has been written, to some degree, for the student audience. In other words, the
details presented here should be sufficient for students to develop their own cycle simulations.
If additional assistance is needed, most of the features of the cycle simulation developed here
have been well documented elsewhere (see, for example, [210]).
International Journal of Mechanical Engineering Education Vol 30 No 4
Instructional version of a thermodynamic cycle simulation 285
The major assumptions and approximations used in the development include the following:
(1) The thermodynamic system is the cylinder contents which are assumed to be spatially
homogeneous, of constant composition, and to obey the ideal gas equation of state;
(2) the specific heat capacities (and the ratio of specific heat capacities) are constant;
(3) most of the other thermodynamic properties (including pressure and temperature) vary
only with time (crank angle) and are spatially uniform;
(4) the flow rates are determined from quasi-steady, one-dimensional flow equations;
(5) the intake and exhaust manifolds are assumed to be infinite plenums containing gases at
constant temperature and pressure;
(6) the fuel is assumed to be completely vaporized and mixed with the incoming air;
(7) the combustion efficiency was assumed to be 100% (i.e. no unburnt fuel); and
(8) the blow-by was assumed to be zero.
Fig. 1 is a schematic of the engine cylinder which shows cylinder heat transfer, work, in-
take flows, exhaust flows, and combustion. The first law of thermodynamics on a differential
rate basis for this system is
Fig. 1. Schematic of the thermodynamic system.
d
d d
d
d
net
in in out out
( )

mu Q
p
V
m h m h

_
,

+ (1)
where m is the total mass in the cylinder, u is the internal energy, is the crank angle, Q
net
is the
net energy (heat) input (energy) release from combustion
*
plus energy gain due to heat transfer
to the cylinder gas), p is the cylinder pressure, V is the cylinder volume,

m
in
is the total mass
flow into the system, h
in
is the specific enthalpy into the system,

m
out
is the total mass flow out
of the system, and h
out
is the specific enthalpy out of the system.
Since the internal energy is only a function of temperature (i.e. no dependence on pressure
or composition) [710], this allows equation (1) to be rewritten as
* The use of the energy (heat) input due to combustion is predicated on the use of enthalpy values which
only consist of the sensible component and do not contain the chemical component.
International Journal of Mechanical Engineering Education Vol 30 No 4
286 Jerald A. Caton

d
d d
d
d
net
in in out out
p
V
Q
p
V
m h m h

_
,

_
,

'

( )

1
1
(2)
To solve equation (2), a number of items are needed. These include the total net energy
(heat) input, the cylinder volume change, and the mass flow rates. The rate of total energy
(heat) input is given by
d
d
d
d
d
d
net app heat
Q Q Q

_
,

_
,
+

_
,
(3)
where (dQd)
app
is the apparent heat release from combustion, and (dQd)
heat
is the energy
gain to the cylinder gases due to heat transfer. The apparent heat release may be determined
from an expression for the burned fuel mass. A number of such expressions exist [5]; for this
work the sub-model chosen for the burned fuel mass fraction was the cosine law [5]. The
derivative of this burn rate model yields the following apparent heat release rate due to
combustion
d
d
app
Q LHV m
AF
b
o
b

_
,

+

_
,

( )
( )( )
sin
2 1
(4)
where LHV is the lower heating value of the fuel, m is the total charge mass,
o
is the start of
combustion,
b
is the combustion (burn) duration, and AF is the mass air to fuel ratio.
In general, the combustion duration,
b
, is a complex function of turbulence, composition,
chemistry, and other factors. Although the turbulence is known to increase with engine speed,
the increase is not thought to be enough to completely compensate for the shorter available
times at the higher engine speeds. For this work, the combustion duration is approximated by a
basic function of engine speed and equivalence ratio. This relationship was recommended by
Campbell [11] based on experimental data reported by Taylor [12]. The relationship is

b
N
+

_
,
+ 40 5
600
1 166 1 1
2
( . ) (5)
where N is the engine speed (rpm) and is the equivalence ratio. The dependence on engine
speed included assists in accounting for the greater number of crank angles involved in the
combustion process as engine speed increases.
The heat transfer to the cylinder gases is given by
d
d
heat
conv wall
Q
h A T

_
,
( )( ) (6)
where h
conv
is the convective heat transfer coefficient, A() is the total surface area in the
cylinder subject to the convective heat transfer, T
wall
is the cylinder wall temperature, and T is
the instantaneous average cylinder gas temperature. The basic convective heat transfer coeffi-
cient used in this work is from Eichelberg [13], but others would be just as suitable [5]. The
total surface area consists of the cylinder walls, the piston top, and the bottom of the cylinder
head, and varies with crank angle. For this work, the cylinder wall temperature is assumed to
be 450 K.
The remaining major task is to determine the flow rates for use in equation (2). These may
be found from basic relations (see, for example, [5, 9, and 14]) for both subsonic and sonic
International Journal of Mechanical Engineering Education Vol 30 No 4
Instructional version of a thermodynamic cycle simulation 287
flows. These relationships are based on a number of assumptions and approximations which
include that the flow is quasi-steady, one-dimensional, adiabatic and incompressible. Since a
real flow would not conform to the above assumptions and approximations, these considera-
tions are corrected by use of an empirical discharge coefficient [5]. In addition, for this work,
unsteady flow dynamics are included (in an approximate manner) by the use of an empirical
adjustment factor. This factor artificially increases the inlet manifold pressure, and thus, in-
creases the flow rates through the engine [9]. The procedure for selecting this adjustment factor
is described below. Since the valve lift profiles are not necessarily known, this work has used a
standard assumption for the instantaneous valve lift based solely on the maximum lift and the
valve open duration, and using a sinusoidal shape as recommended by Sherman and Blumberg
[14].
To complete the required input information, the boundary conditions for the inlet (tempera-
ture and pressure) and for the exhaust (pressure) are needed. In general, these parameters are
complex functions of the engine design, application and operating conditions. For purposes of
this work, therefore, the inlet temperature is assumed constant. With specific information, how-
ever, the inlet temperature could easily be a function of engine operation. The inlet and exhaust
pressures were assumed to vary with engine speed, and in addition, the exhaust pressure was
assumed to vary with load (inlet pressure) [9].
The instantaneous cylinder pressure as a function of crank angle is obtained by the numeri-
cal integration of the pressure differential equation (equation (2)). At each time step, all param-
eters are recalculated. The numerical method chosen was a basic Euler technique with a
predictor-corrector adjustment. Once the cylinder pressure is determined for each calculation
step, the ideal gas equation of state is used to determine the corresponding average cylinder gas
temperature.
To begin a particular engine cycle calculation, several parameters are not known. These are
the initial amount of exhaust gases (residual), and the initial cylinder gas temperature and
pressure and must be assumed. After 720 crank angles (for a four-stroke cycle engine), the final
values of these three parameters are compared to the initial guesses. If the final values are not
within an acceptable tolerance to the initial values, the calculation is repeated using the final
values for the initial values. All else the same, this procedure usually finds convergence within
about three complete cycles.
Once the cylinder pressures are determined for a given case, the indicated power may be
found by integrating the cylinder pressure as a function of cylinder volume [5]. Once the indi-
cated power is known, the other indicated performance parameters may be determined using
well-established relations [5, 9].
To determine the brake performance values, the engine frictional losses must be estimated.
Since the frictional losses for the specific engine studied here were not known, a generic ex-
pression was chosen to estimate the engine friction [5]:
fmep bar
N N
( ) . . . +

_
,
+

_
,
0 97 0 15
1000
0 05
1000
2
(7)
where fmep is the frictional mean effective pressure (and has units of bar for these constants)
and N is the engine speed (rpm). Although more exact expressions could be used, for purposes
of this work, the expression used provides the basic characteristic of increasing frictional losses
with engine speed.
To obtain the brake values of the performance parameters, the brake mean effective pres-
sure (bmep) is first determined:
bmep = imep fmep (8)
International Journal of Mechanical Engineering Education Vol 30 No 4
288 Jerald A. Caton
where imep is the indicated mean effective pressure. The other brake performance values may
now be determined based on the brake mean effective pressure.
RESULTS AND DISCUSSION
Again, the purpose of this work was not to simply provide yet another set of results from a
standard thermodynamic cycle simulation, but rather, to provide examples of the type of results
that these relatively simple instructional versions of cycle simulations are able to yield. These
examples are not comprehensive, but are intended to illustrate the relatively detailed and
relevant results that are possible.
Engine specifications and operating conditions
To illustrate the use of the above cycle simulation, a modern, commercial, spark-ignition en-
gine was selected. The engine is a naturally aspirated, 60 V-6 configuration, with port fuel-
injection, four valves per cylinder, and two overhead cams [15]. Table 1 outlines the key
specifications and features of this engine.
Table 1. Specifications for the engine.
Parameter Value
Engine type
Engine construction
Bore stroke
Displacement
Geometric comp. ratio
Valve diameter: intake
Valve diameter: exhaust
Maximum valve lift: (both intake and exhaust)
Intake valve opening
Intake valve closing
Exhaust valve opening
Exhaust valve closing
Design power at rpm
Design torque at rpm
V-6, 60 cylinder banks
Cast iron block, aluminium cylinder
heads
92 84 mm (3.62 3.31 in)
3.35 litres (206 in
3
)
9.25:1
36.5 mm
32.0 mm
9.4 mm
342aTDC
110aTDC (610aTDC)
130aTDC
388aTDC
149 kW at 5000 rpm
(200 hp at 5000 rpm)
291 Nm at 4000 rpm
(215 ftlb at 4000 rpm)
Basic performance results
With the above inputs, the engine performance was computed as a function of engine speed.
For each engine speed, the spark timing was adjusted to determine the maximum brake torque
Although most of the key engine specifications were available from the open literature
[15], the remaining specifications were assumed, and Table 2 lists these values. As demon-
strated elsewhere [9], for the most part, the performance results are not extremely sensitive to
the parameters which required assumed values.
International Journal of Mechanical Engineering Education Vol 30 No 4
Instructional version of a thermodynamic cycle simulation 289
Table 2. Values of selected parameters.
Parameter Value
Fuel
Fuel heating value
Airfuel ratio (stoichiometric)
Inlet air temperature
Cylinder wall temperature
Exhaust pressure
Inlet pressure
Rod length to crank radius
Valve discharge coefficient
Specific heat capacity ratio
Specific gas constant
Ignition delay
Gasoline
42280 kJkg
14.67
313 K
450 K
Variable [9]
Variable [9]
4.44
0.70
1.35
0.287 kJkgK
20
(MBT) for the given operating conditions (variation of performance with spark timing is de-
scribed below). Unless stated otherwise, the following results represent engine performance
for MBT spark timing.
The procedure used to generate computed engine performance as a function of engine speed
involved several steps. First, the uncorrected brake torque was determined as a function of
engine speed. For low speeds (where the flow dynamics would be expected to be less impor-
tant), these values were in reasonably good agreement with the measured values. For the mid-
speed range (say 4000 to 6000 rpm) the agreement was not acceptable. To compensate for the
discrepancy, an adjustment factor was determined which provided good agreement. This
adjustment factor increased the intake pressure by an artificial amount to increase the flow
which resulted in increase torque and power. Although this adjustment is assumed mostly
necessary to compensate for the lack of unsteady flow dynamics in the simulation, this adjust-
ment also accounts for any adjustment needed for the other sub-models such as combustion,
friction, heat transfer, and so forth.
The adjustment factor was determined at wide open throttle conditions, and then assumed
to be valid for part load operation. This assumption has been used elsewhere (e.g., see Miller et
al. [16]). This implies that the manifold unsteady flow dynamics are approximately constant
as a function of load, but vary with engine speed. This is consistent with the notion that the
flow dynamics are dominated by overall flow changes which are proportional to engine
speed.
Fig. 2 shows the experimental (solid curves) [15] and computed (dashed lines with
symbols) brake power and torque as functions of engine speed. Also shown, is the adjustment
factor used to compensate for the unsteady flow dynamics. These values were found by
examination. This adjustment factor accounts largely for the lack of detail on the unsteady flow
dynamics, but also includes adjustments for any of the inaccuracies associated with other
portions of the cycle simulation. In general, the results of the computations are in good agree-
ment with the experimental values. Based on this modest verification, other details of the
engine performance were computed and are discussed next. First, the results for instantaneous
quantities as a function of crank angle (time-resolved) for the base case will be presented.
Then, the results for the overall engine performance are presented.
International Journal of Mechanical Engineering Education Vol 30 No 4
290 Jerald A. Caton
Fig. 2. Experimental (solid curves) [15] and computed (dashed lines with symbols) brake
power and torque as functions of engine speed. Also the adjustment factor used to compensate
for the unsteady flow dynamics as a function of engine speed.
Time-resolved results
Fig. 3 shows the cylinder pressure and average gas temperature as a function of crank angle for
the 5000 rpm case at wide open throttle. As shown, the cylinder pressure peaks at about
12.5aTDC with a value of about 6500 kPa, and the average cylinder gas temperature peaks at
about 34aTDC with a value of about 2950 K. Since these results were generated for constant
specific heat capacities, the computed values for the maximum cylinder pressures and tem-
peratures are expected to be higher than the actual values (actual cylinder pressures were not
available).
Fig. 4 shows the total exhaust and intake mass flow rates, and the cylinder pressure for the
5000 rpm case for wide open throttle. For this figure, the positive flow rates reflect flow in the
normal directions (i.e. the exhaust flow is positive out of the cylinder and the intake flow is
positive into the cylinder). Further, for this figure and the following discussion, the intake and
exhaust flow rates represent the total from both passages for the intake and exhaust systems,
respectively. Recall that this engine possesses two valves and passages for both the intake and
exhaust systems.
As shown in Fig. 4, the exhaust flow increases rapidly once the exhaust valve begins to
open at 130aTDC. The exhaust mass flow reaches a maximum value of about 0.21 kgs at
155aTDC and then begins to decrease. The exhaust flow rate reverses (the flow is from the
exhaust system into the cylinder) at about 345aTDC until the exhaust valve closes. When the
intake valve opens at 342aTDC, the pressure in the cylinder is greater than the manifold
pressure. This results in flow from the cylinder into the intake manifold. Eventually, the pres-
sure in the cylinder becomes lower than the intake manifold pressure so that flow can begin to
enter the cylinder. The intake flow reaches a maximum flow rate of about 1.8 kgs and then
decreases.
International Journal of Mechanical Engineering Education Vol 30 No 4
Instructional version of a thermodynamic cycle simulation 291
Fig. 3. Cylinder pressure and temperature as a function of crank angle for the 5000 rpm case
at wide open throttle.
Fig. 4. Exhaust and intake mass flow rates, and cylinder pressure for the 5000 rpm case for
wide open throttle.
Overall performance results
Fig. 5 shows the brake torque as functions of spark timing for five different engine speeds. As
mentioned above, the spark timings are relative since the ignition delay was not known and
was assumed constant (20) for all cases. Even for these relative spark timings, however, the
International Journal of Mechanical Engineering Education Vol 30 No 4
292 Jerald A. Caton
Fig. 5. Computed brake torque as functions of spark timing for five different engine speeds
for wide open throttle.
trends are valid. Furthermore, the crank angles for peak pressure and temperature will still be
accurate, and will reflect the actual mechanics and thermodynamics.
As shown in Fig. 5, for each engine speed, an optimum (MBT) spark timing exists which
maximizes the brake torque. Note that the higher the engine speed, the more advanced the
spark timing must be to achieve maximum brake torque. These values and trends are consistent
with published values and trends for similar automotive engines [5].
Fig. 6 shows the volumetric, mechanical, indicated thermal, and brake thermal efficiencies
as functions of engine speed for wide open throttle. These efficiencies are optimistic values
since the fuel conversion was assumed to be 100%, and the blow-by flow was assumed to be
zero. With additional information, these latter two assumptions could be relaxed and more
accurate values would be obtained for these efficiencies.
As shown in Fig. 6, the volumetric efficiency
*
is about 86% at engine speeds less than 2000
rpm and is as high as 100% for an engine speed of 5000 rpm. The variation of volumetric
efficiency as a function of engine speed reflects the values of the empirical adjustment factor
which increases the flow rates to the engine to provide greater agreement with the reported
brake torque. In general, the functional dependence of the volumetric efficiency with respect to
engine speed will be similar to the corresponding dependencies of the torque and brake mean
effective pressure.
The mechanical efficiency shown in Fig. 6 is about 88% at 1000 rpm and steadily decreases
to about 62% at 7000 rpm. This result of the mechanical efficiency reflects the increasing
importance of the friction as engine speed increases. The indicated thermal efficiency varies
between about 36 and 42%, and the brake thermal efficiency varies between about 32 and 35%
for these engine speeds and operating conditions. The brake thermal efficiency decreases more
rapidly for increasing engine speeds due to the increasing value of the mechanical friction.
*
Note that volumetric efficiency may be defined in a number of ways [5]. The definition used here is
described in [9].
International Journal of Mechanical Engineering Education Vol 30 No 4
Instructional version of a thermodynamic cycle simulation 293
Fig. 6. Volumetric, mechanical, indicated thermal, and brake thermal efficiencies as functions
of engine speed for wide open throttle.
Fig. 7 shows the computed brake mean effective pressure (bmep), brake specific fuel con-
sumption (bsfc), brake thermal efficiency, and the percentage heat loss as a function of engine
speed for wide open throttle. As shown, the bmep is proportional to the computed torque (see
Fig. 2). The specific fuel consumption and thermal efficiency curves are inverses of each
Fig. 7. Computed brake mean effective pressure, brake specific fuel consumption, brake
thermal efficiency, and the percentage heat loss as a function of engine speed for wide open
throttle.
International Journal of Mechanical Engineering Education Vol 30 No 4
294 Jerald A. Caton
other, and show a maximum fuel efficiency for an engine speed of about 4000 rpm. Finally, the
percentage heat loss (which is the amount of cylinder heat loss as a fraction of the fuel energy)
decreases from about 38% at 1000 rpm to about 17% at 7000 rpm. This decrease of the percent-
age heat loss with increasing engine speed is largely due to the shorter available real times
available as the engine speed increases. Again, these results are consistent with reported results
for similar automotive engines [5].
Fig. 8 shows peak cylinder pressure, crank angle of peak pressure and combustion duration
as functions of engine speed for wide open throttle for MBT spark timings. The results for this
figure were obtained from a sequence of figures like Fig. 5 for each engine speed. For these
conditions, the peak cylinder pressure ranges between about 5.2 and 6.5 MPa. The crank angle
of the peak cylinder pressure varies between about 19aTDC for 1000 rpm to about 10aTDC
for 7000 rpm. The crank angle of the peak cylinder pressure is largely dependent on the com-
bustion duration. The combustion duration increases from about 45 at 1000 rpm to 95 at
7000. The variation of the crank angle for the peak pressure is largely due to the variation of the
combustion duration as a function of engine speed. Also, shown on this figure are values for the
crank angle of peak cylinder pressure for a different engine with different operating conditions
at part load but with similar combustion durations (5060) from [14]. The agreement be-
tween the two studies is remarkable, particularly since the engine and operating conditions
were different.
Fig. 8. Peak cylinder pressure, crank angle of peak pressure and combustion duration as
functions of engine speed for wide open throttle for MBT spark timings. Also, this figure
shows values for the crank angle of peak cylinder pressure for a different engine with different
operating conditions (see text) from [2].
Fig. 9 shows the peak cylinder temperature, mass-average exhaust temperature, and the
crank angle of the peak temperature as functions of engine speed for wide open throttle. The
peak cylinder temperature is about 2990 K for 1000 rpm, and about 2880 K for 7000 rpm. The
crank angle for the peak cylinder temperature is about 27.6aTDC for 1000 rpm, and is about
34.5aTDC for 7000 rpm. This variation of the crank angle of the peak temperature with
engine speed reflects the longer combustion duration with increasing speed (see Fig. 8).
International Journal of Mechanical Engineering Education Vol 30 No 4
Instructional version of a thermodynamic cycle simulation 295
Fig. 9. Peak cylinder temperature, mass-average exhaust temperature, and the crank angle
of the peak temperature as functions of engine speed for wide open throttle.
The mass-averaged exhaust temperature [9] is about 1050 K for 1000 rpm, and increases to
about 1400 K for 7000 rpm. This increasing exhaust temperature reflects the decreasing per-
centage of cylinder heat loss at higher engine speeds (see Fig. 7). As more of the fuel energy is
retained by the cylinder gases at higher engine speeds, the thermal efficiencies decrease slightly
and the exhaust energy increases.
The above results have been for wide open throttle (WOT) engine operation. In contrast, the
next results are examples of part load performance. Fig. 10 shows the brake mean effective
pressure as functions of inlet pressure for four engine speeds. As shown, the brake mean
effective pressure ranges from less than 100 kPa to over 1000 kPa for the engine operating
conditions considered in this paper. The increase in brake mean effective pressure is almost
linear with inlet pressure for each engine speed. Note that the maximum brake mean effective
pressure for each engine speed represents the performance at wide open throttle. These results
are similar to those for typical automotive engines [5].
SUMMARY AND CONCLUSIONS
This paper described the development and use of an instructional version of a thermodynamic
cycle simulation for use by undergraduate students. To achieve this goal, the recommended
simulation was based on fundamental features, and with constant specific heat capacities (and
the associated constant composition). In spite of these limitations, the instructional version of
the cycle simulation was able to capture a number of significant engine characteristics.
To illustrate the usefulness of such a basic simulation, this simulation was used to explore
the overall performance of a commercial automotive spark-ignition engine for a range of
engine speeds and operating conditions. Although the cycle simulation was based on
elementary conventional features, a number of important engine characteristics were correctly
obtained. These engine characteristics included the power, torque, mean effective pressure,
International Journal of Mechanical Engineering Education Vol 30 No 4
296 Jerald A. Caton
specific fuel consumption, thermal efficiency, heat loss, volumetric efficiency, and mechanical
efficiency as functions of engine loads and speeds (up to 7000 rpm).
Fig. 10. Brake mean effective pressure as functions of inlet pressure for four engine speeds.
REFERENCES
[1] McAulay, K. J., Wu, T., Chen, S. K., Borman, G. L., Myers, P. S., and Uyehara, O. A., Development
and evaluation of the simulation of the compression-ignition engine, Society of Automotive
Engineers, SAE Paper No. 840451, 1965.
[2] Heywood, J. B., Higgins, J. M., Watts, P. A., and Tabaczynski, R. J., Development and use of a cycle
simulation to predict SI engine efficiency and NO
x
emissions, Society of Automotive Engineers,
SAE Paper No. 790291, 1979.
[3] Heywood, J. B., Engine Combustion Modellingan Overview, in Combustion Modelling in
Reciprocating Engines, eds., J. N. Mattavi and C. A. Amann, Plenum Press, New York, NY, 1980.
[4] Sorenson, S. C., Simple computer simulations for internal combustion engine instruction, Inter-
national Journal of Mechanical Engineering Education, 9(3), 237249, 1981.
[5] Heywood, John B., Internal Combustion Engine Fundamentals, McGraw-Hill Book Company, New
York, 1988.
[6] Ramos, J. I., Internal Combustion Engine Modelling, Hemisphere Publishing Corporation, New York,
NY, 1989.
[7] Caton, J. A., A comparison of the use of constant or variable properties in a thermodynamic cycle
simulation for a spark-ignition engine, Proceedings of the 1999 Joint Meeting of the United States
Sections: The Combustion Institute, Paper No. 99-49, pp. 196199, The George Washington
University, Washington, DC, 1517 March 1999.
[8] Caton, J. A., Comparisons of instructional and complete versions of thermodynamic engine cycle
simulations for spark-ignition engines, accepted for publication, International Journal of
Mechanical Engineering Education, 29(4), 283306, 2001.
[9] Caton, J. A., A basic thermodynamic engine cycle simulation: complete description and sample
application to a commercial automotive spark-ignition engine, Report No. ERL-98-01, Engine
Research Laboratory, Texas A&M University, Department of Mechanical Engineering, Version 1.0,
22 March 1998.
International Journal of Mechanical Engineering Education Vol 30 No 4
Instructional version of a thermodynamic cycle simulation 297
[10] Caton, J. A., Development and use of a basic thermodynamic cycle simulation for commercial auto-
motive spark-ignition engines, Proceedings of the 1998 Central States Section/Combustion Institute
Spring Technical Meeting, Paper No. 98-06, pp. 2934, Marriott Griffin Gate Hotel, Lexington, KY,
31 May2 June 1998.
[11] Campbell, A. S., Thermodynamic Analysis of Combustion Engines, John Wiley & Sons, New York,
1979.
[12] Taylor, C. F., The Internal Combustion Engine in Theory and Practice, Vol. II, Chapter 1, MIT Press,
Cambridge, MA, 1968.
[13] Eichelberg, G., Some new investigations on old combustion engine problems, Engineering, 148,
p. 463, 1939.
[14] Sherman, R. H., and Blumberg, P. N., The influence of induction and exhaust processes on
emissions and fuel consumption in the spark ignited engine, SAE Paper No. 770990, 1977.
[15] Subhedar, Jay W., Leskiw, P. M., Hahn, C. J., and Langdon, T. P., General Motors 3.4L Twin Dual
Cam V6 engine, SAE Paper No. 910679, 1991.
[16] Miller, R. H., Lavoie, G. A., Gardner, T., Davis, G. C., and Newman, C. E., A super-extended
Zeldovich mechanism of NO
x
modeling and engine calibration, presented and published at the
1998 SAE Congress, SAE Paper No. 980781, February 1998.

Das könnte Ihnen auch gefallen