Sie sind auf Seite 1von 14

SAE TECHNICAL

PAPER SERIES

1999-01-1665

Modeling of Engine Exhaust Acoustics


Thomas Morel and John Silvestri
Gamma Technologies,

Karl-Alfred Goerg
BMW AG

Rolf Jebasinski
J. Eberspacher, GmbH & Co.

Reprinted From: Proceedings of the 1999 Noise and Vibration Conference


(P-342)

Noise and Vibration Conference & Exposition


Traverse City, Michigan
May 17-20, 1999
400 Commonwealth Drive, Warrendale, PA 15096-0001 U.S.A.

Tel: (724) 776-4841 Fax: (724) 776-5760

The appearance of this ISSN code at the bottom of this page indicates SAEs 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 Sections 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 1999 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

1999-01-1665

Modeling of Engine Exhaust Acoustics


Thomas Morel and John Silvestri
Gamma Technologies,

Karl-Alfred Goerg
BMW AG

Rolf Jebasinski
J. Eberspacher, GmbH & Co.
Copyright 1999 Society of Automotive Engineers, Inc.

ABSTRACT

these two design targets are difficult to optimize simultaneously by standard hardware testing procedures, simulations are called on to assist in this process. In recent
years, much work has been done on improving the predictive capabilities for intake/exhaust acoustics, to the
point where they are today serving as useful tools for
design and optimization. These efforts have been well
documented in the literature, e.g. by Torregrosa et al [2],
Selamet et al [3], Ferrari and Onorati [4], Pilo et al [5],
Challen [6] and Morel et al [7], showing the increasing
usefulness of simulations in the acoustics design process.

Exhaust acoustics simulation is an important part of the


exhaust system process. Especially important is the
trend towards a coupled approach to performance and
acoustics design. The present paper describes a new
simulation tool developed for such coupled simulations.
This tool is based on a one-dimensional fluid dynamics
solution of the flow in the engine manifolds and exhaust
and intake elements. To represent the often complex
geometries of mufflers, an easy-to-use graphical preprocessor is provided, with which the user builds a model
representation of mufflers using a library of basic elements. A comparison made to two engines equipped
with exhaust silencers, shows that the predictions give
good results.

The objective of this work has been the application of a


new CAE tool (GT-Power) designed for simultaneous
engine performance and acoustics analysis. GT-Power is
a detailed engine simulation tool, specifically focused on
transient engine processes, and containing detailed models of engine breathing (with and without turbocharging),
combustion, emissions, and thermal warm-up of components. To aid in the development of exhaust system models, GT-Power contains a dedicated muffler preprocessor
tool which allows the user to build a model graphically by
point-and-click procedures. This new tool was tested by
comparison to experimental data obtained on two
engines equipped with exhaust silencers.

INTRODUCTION
The engine and vehicle development process is subject
to steadily increasing customer expectations, including
requirements regarding intake and exhaust noise. In the
past, the design process was split into two parts, first
concentrating on the issues of engine performance
alone, with emphasis on good breathing and high volumetric efficiency (and engine torque). Only as the second step, the process moved on to the design of the
intake and exhaust elements to control and reduce the
intake and exhaust noise. It has since been realized that
the design modifications implemented concerning the
acoustics could be in conflict with the performance objectives accomplished in the first step of the development.
Furthermore, some of the engine design parameters, e.g.
valve profiles and timing, selected in the performance
development can quite profoundly affect the system
acoustics (Silvestri et al. [1]).

ENGINE SYSTEM MODEL


GT-Power is a comprehensive engine performance analysis code (Morel et. al [8]), which contains highly detailed
models of all key engine processes, as discussed below.
This tool is a part of GT-SUITE, which is an integrated set
of CAE tools for analysis of engines and powertrains.
ENGINE BREATHING The flow solution is carried out
by the solution of one-dimensional flow equations by a
finite difference scheme. The entire system is discretized
into many small control volumes on which the governing
equations are evaluated on a staggered mesh. This

As a consequence of the recognition of these mutual


effects, the trend in the vehicle industry is now toward
integrated engine performance and acoustics. Since
1

means that all scalars (pressure, temperature, species)


are calculated at the center of the volumes, while the vectors (velocity, mass flux and enthalpy flux) are calculated
at the volume boundaries. The main flow elements are
pipes (which are sub-divided into a number of subvolumes) and flowsplits (volumes which have more than
two openings and which can represent dividing streams,
volumes separated by perforated walls and other complex flow domains). The system representation is built
from various components (pipes, flowsplits, cylinders,
crankcases, compressors, turbines, burners, etc.) which
are connected by connections (orifices, valves, throttles, CFD connections), as shown in Figure 1. This figure
illustrates a 6-cylinder truck diesel engine.

Design of exhaust systems often covers emission control


equipment in addition to the acoustics. One of the features of the solution is the calculation and tracking of all
of the species of products of combustion. This allows the
prediction of emissions and also provides the necessary
boundary conditions for catalyst chemistry calculations.
Work on these emissions calculations is in progress.
THERMAL SOLUTION AND ENGINE WARM-UP
Exhaust gas temperature has a large effect on speed of
sound and thus on the acoustics. Consequently, an
important aspect of engine operation is heat transfer (incylinder and in intake/exhaust manifolds) and resulting
component temperatures. Heat transfer is a strong function of engine speed and load, and so are the component
temperatures. Under transient engine operation there is
also the effect of thermal inertia, which has important
effect on a variety of diverse issues such as turbocharger
response, catalyst light-off and knock. To model these
one has to calculate the heat transfer coefficients
between the gases and the walls, and also solve the heat
conduction in the solid structures (pistons, pipes, etc.).

An optimization of the breathing process can be carried


out using a special general procedure which allows the
automatic optimization of practically any input parameter
with respect to any output parameter. This is illustrated in
Figure 2, showing the automatic optimization of the ziptube length for maximum volumetric efficiency.

The in-cylinder heat transfer model uses several alternative ways to calculate the heat transfer coefficients. One
is the Woschni [9] model, widely used in the industry,
which uses the mean piston speed for Reynolds number
calculations. A more accurate calculation uses an incylinder flow model to calculate the instantaneous flow
velocities (swirl, squish and turbulence are used for Reynolds number calculation) to get spatially resolved heat
transfer coefficients (Morel and Keribar [10]. Other
options include importing data from CFD calculations and
links to user-supplied subroutines. The heat conduction
solution is FE based, where FE models of in-cylinder
components (piston, rings, liner, head, valves) are built
automatically using generic templates. A built-in FE
solver calculates steady-state or transient temperature
distributions including the effect of temperature dependent thermal properties.

Figure 1. Model of a 6-cylinder, turbocharged, turbocompound diesel engine.

The manifold heat transfer calculations use a different


approach. The heat transfer coefficient is based on the
Colburn analogy, using the instantaneous flow velocity in
the pipes for the Reynolds number calculation. The conduction through the pipe walls assumes axial symmetry,
and it allows the walls to consist of many layers of different materials. Air layers (air-gaps) are also allowed. The
solution accounts both for radial and axial flow of heat
through the pipe walls. During transient operation, this
model tracks the warm up of the manifolds and pipes,
including the warm up of the catalyst.
COMBUSTION AND EMISSIONS Several combustion
models are available for spark ignited and diesel engines.
Most often used are three simple models: user-specified
combustion profiles and Wiebe function models for S.I.
and diesel engines. In addition, there are two detailed
combustion models. For diesel engines there is a diesel
jet combustion model, which represents an injected jet/
plume by up to 1500 subvolumes; for homogeneous

Figure 2. Results of automatic determination of optimal


zip tube length for maximum volumetric
efficiency, using generic optimizer, in a 4cylinder automotive engine.

charge spark ignition engines there is a turbulent flame


front model (Morel et al [8]).

ent configurations were studied. One was the baseline


system with a straight pipe, made of several pipe elements, and no silencer. The other two were based on
this baseline, replacing pieces of the straight pipe by
silencing elements. In one of these cases the silencing
element was a concentric tube resonator, in the other it
was a simple three-pass muffler.

VEHICLE TRANSIENTS Although GT-Power is nominally a thermodynamics cycle model, it in fact has significant cranktrain dynamics and vehicle capabilities (Morel
et al, 1999). The torque and speed calculations are carried out by a single-degree-of-freedom dynamics model
of all of the moving parts, including: pistons, connecting
rods, crankshaft and flywheel. Cylinder and crankcase
pressures act on the piston from top and bottom to generate instantaneous torque at the flywheel. Engine
speed fluctuations during each engine cycle are also calculated. If needed, users can attach a vehicle to the
engine and calculate, for example, a vehicle acceleration
complete with transmission gear shifting. This capability
is useful in transient vehicle noise analysis.

The baseline system was a straight pipe, of total length of


3455 mm. There were 2 transducers. One was installed
at an upstream location 2690 mm from the tail end of the
pipe (location 1). The other was near the exit (tailpipe),
420 mm from the end (location 7).
The concentric tube resonator was made of a cylindrical
shell 128 mm diameter and 380 mm long. The center
pipe had 846 holes 3.5 mm in diameter, arranged in two
patches (Figure 4). This resonator was inserted into the
exhaust system, replacing a pipe of the same length.
There were 4 transducers. One pressure transducer was
installed at the upstream location 1, one inside the resonator shell, one just downstream of the resonator (location 2), and one near the tail pipe (location 7).

MUFFLER PREPROCESSOR
A graphical pre-processor has been developed for building models of exhaust components. This pre-processor
is built directly into the GT-SUITE graphical interface. It
contains a library of elements, from which one can build
up a model of a muffler. These elements include: muffler
shell, straight pipe, bent pipe, baffle, perforations for
pipes, perforations for baffles and concentric tube resonators. The volumes inside the shell can be partly filled
with absorptive acoustic material (e.g. mineral wool).
The user builds the model graphically by dragging and
sizing the basic elements (Figure 3).

Figure 4. Concentric tube resonator generated by the


muffler preprocessor.
The muffler was a basic three-pass design with three
pipes and two baffles, with perforations in one of the baffles (Figure 5). This resonator was inserted into the
exhaust system, replacing a pipe of the same length.
Comparisons were made for four pressure transducers.
Of these, one pressure transducer was installed at the
upstream location 1, the second just upstream of the
muffler (location 2), the third inside the muffler in the volume between the second baffle and the muffler end, and
the fourth one near the tail pipe (location 7).

Figure 3. Typical tri-flow muffler design generated by the


muffler graphical preprocessor

APPLICATION TO 6-CYLINDER ENGINE WITH


TWO DIFFERENT REACTIVE SILENCERS

RESULTS FOR THE STRAIGHT PIPE BASELINE The


results for the straight pipe are shown in Figures 6 a, b
(6000 rpm), 6c, d (4500 rpm) and 6 e, f (3000 rpm).
Since there is no muffler, the pressure magnitude is
about the same at both locations. It is seen that the simulation well reproduces the pressures at both locations,
and their changing characteristics from speed to speed.

An experimental data set was acquired on a specially


designed engine setup, based on a BMW 6-cylinder inline passenger car engine. The exhaust system was
arranged into two parallel parts, each connected to one
bank of the engine. Of these, one side was instrumented
with pressure transducers, measuring instantaneous
pressures along the length of the exhaust. Three differ3

Figure 5. Three-pass muffler generated by the muffler


preprocessor.
RESULTS FOR CONCENTRIC TUBE RESONATOR
For this case, the results are shown in Figures 7 (6000
rpm) and in Figure 8 (3000 rpm). The measured pressures show the attenuation produced by the resonator,
which can be seen by comparing the first location with
the three other locations. Again, the simulation reproduces reasonably well the measured values.
RESULTS FOR MUFFLER The results for the threepass muffler are shown in Figure 9 (6000 rpm) and Figure 10 (3000 rpm). They show strong attenuation, especially at the last two locations, as compared to the two
upstream locations. Again, the simulation agrees well
with the measurement at all locations.

APPLICATION TO A V-6 ENGINE WITH AN


ABSORPTIVE SILENCER
Another experimental data set was acquired on a V-6
Audi engine, equipped with a single exhaust silencer.
This silencer was a concentric tube resonator, made of a
cylindrical shell 156 mm diameter and 480 mm long. The
center pipe had 2280 holes 3.5 mm in diameter, uniformly
distributed along the entire length of the resonator (Figure 11). There were two transducers, one installed at an
upstream location 180 mm ahead of the resonator, the
second 180 mm downstream of the resonator. The
experiments were run over a wide range of speeds, from
1000 rpm to 6500 rpm. There were two sets of data, one
with an empty resonator, one with the resonator filled with
absorbing wool material, packed with a density of 120 g/
liter. The data taken at the two locations were presented
in the form of pressure level at various orders across the
wide speed range.
RESULTS FOR EMPTY RESONATOR Ahead of the
resonator (Figure 12) the simulations gave a good agreement with the measurements. A similar level of agreement was found at the location after the resonator, except
for the 3rd order at the higher speeds (Figure 13).

Figure 6. a,d. Straight pipe, 6000 and 4500 rpm.

CONCLUSIONS
1. An engine system simulation model has been developed, with applications to simultaneous performance
and acoustics engine design.
2. Comparisons to engine data with several different
exhaust mufflers showed that the simulated exhaust
pressure dynamics closely match the experimental measurements in both frequency content and amplitude.
3. The model is applicable to both reactive and absorptive (with mineral wool) exhaust components.

ACKNOWLEDGMENTS
Thanks are due to Mr. Bernard Challen of Shoreham Services for his assistance during the model development.
His practical and theoretical experience led to many valuable suggestions along the way.

REFERENCES
1. J.J. Silvestri, T. Morel and M. Costello, "Study of
Intake System Wave Dynamics and Acoustics by
Simulation and Experiment", SAE Paper 940206,
February 1994.
2. A.J. Torregrosa, A. Broatch and F. GonzalesContreras, A Theoretical and Experimental Study of
the Behavior of Concentric Perforated Duct Automotive Mufflers, SAE Paper 960300, February 1996.
3. A. Selamet, S.H. Yonak, J.M. Novak, M. Khan, The
Effect of Vehicle Exhaust System Components on
Flow Losses and Noise in Firing Spark-Ignition
Engines, SAE Paper 951260.
4. G. Ferrari and A. Onorati, A Comprehensive 1-D Model
for the Simulation of Gas Flow Through I.C. Engine
Pipe Systems with Chemical Species Tracking, FISITA
Paper F98T154, 1998 FISITA Congress, Paris.
5. L. Pilo, F. Gamba and B.J. Challen, Prediction of
Vehicle Radiated Noise, SAE Noise and Vibration
Conference, Paper 97NV20, May 1997.
6. B.J. Challen, Modern Modelling for I.C. Engines
Intake and Exhaust System Design, Acoustics Bulletin, January/February 1998, pp 5-9.
7. T. Morel, J. Morel and D.A. Blaser, "Fluid-Dynamic
and Acoustic Modeling of Concentric-Tube Resonators/Silencers", SAE Paper 910072, February, 1991.
8. T. Morel, R. Keribar, J.J. Silvestri and S. Wahiduzzaman, Integrated Engine/Vehicle Simulation and Control, SAE Paper 1999-01-0907, March 1999.
9. G. Woschni, [1967], An Universally Applicable Equation for the Instantaneous Heat Transfer Coefficient in
the Internal Combustion Engine, SAE Transactions,
Vol. 76, p. 3065, 1967.
10. T. Morel and R. Keribar [1985], "A Model for Predicting Spatially and Time Resolved Convective Heat
Transfer in Bowl-in-Piston Combustion Chambers",
SAE Paper 850204, SAE Congress, February, 1985.

Figure 6. e,f. Straight pipe, 3000 rpm.


RESULTS
FOR
RESONATOR
FILLED
WITH
ABSORBING WOOL With the wool packing inside the
resonator, the level of agreement in front of the resonator
was about the same as without the wool (Figure 14).
After the resonator the agreement was good for all orders
except the 9th order above 3000 rpm (Figure 15).
EFFECTS OF ABSORBING WOOL The effects of the
absorbing wool are clearly seen by comparing the experimental results from the empty and filled resonators.
There is a significant difference even upstream ahead of
the resonator (Figure 16). This difference is the most
apparent at low speeds in the third order, where the dip in
the pressure level is seen to have moved to lower speeds
due to the presence of the wool packing. Downstream of
the resonator (Figure 17), the 3rd order noise level is
actually higher with the wool, and there are significant
changes up and down at the higher orders. The calculated changes produced by the wool agree very well with
the experiments ahead of the resonator (Figure 18). Also
after the resonator, the main trends produced by the wool
were well predicted (Figure 19).

Figure 8. Concentric tube resonator, 3000 rpm


Figure 7. Concentric tube resonator, 6000 rpm
6

Figure 9. Three-pass muffler, 6000 rpm.

Figure 10. Three-pass muffler, 3000 rpm.

Figure 11. Six-cylinder engine with a concentric tube


resonator.

Figure 12. Empty resonator, upstream location.

Figure 14. Resonator with wool packing, upstream


location.
Figure 13. Empty resonator, downstream location.
9

Figure 15. Resonator with wool, downstream location.


Figure 16. Effect of wool packing, experiment, upstream.
10

Figure 18. Effect of wool packing, calculation, upstream.

Figure 17. Effect of wool packing, experiment, upstream.


11

Figure 19. Effect of wool packing, calculation,


downstream.
12

Das könnte Ihnen auch gefallen