Vous êtes sur la page 1sur 25

1

2006-01-3654
Improvement of Intake Restrictor Performance for a Formula
SAE Race Car through 1D & Coupled 1D/3D Analysis Methods
Mark Claywell and Donald Horkheimer
University of Minnesota


Copyright 2006 SAE International
ABSTRACT
A typical means of limiting the peak power output of race
car engines is to restrict the maximum mass flow of air
to the engine. The Formula SAE sanctioning body
requires the use of an intake restrictor to limit
performance, keep costs low, and maintain a safe racing
experience. The intake restrictor poses a challenge to
improving engine performance. Methods to better
understand the ramifications of the restrictor on the
engine lead to performance improvements that allow an
edge over the competition.
A one-dimensional gas exchange simulation code
coupled with three-dimensional CFD is used to simulate
various concepts in the improvement of restrictor
performance. Ricardos WAVE and VECTIS are the
respective simulation codes. Along with this, the
interaction of intake manifold and restrictor are
considered. The effects of different diffuser geometries
and plenum dimensions were first explored using
WAVE, and then a series of different diffuser angles
were simulated using WAVE-VECTIS. Primary area of
improvement was determined to be through the use of
tighter or smaller diffuser angles.
Acoustic filtering using Helmholtz resonators was
investigated using WAVE to determine if restrictor
performance could be improved by attempting to make
flow at the throat more uniform over the cycle. Inline
Helmholtz resonators were also investigated in an
attempt to increase pressure upstream of the throat.
Initial results were encouraging, but were very sensitive
to geometry. An additional coupled simulation
considered the effect of swirl vanes placed upstream of
the restrictor throat. Swirl vanes had little to no effect on
the performance of the intake.
INTRODUCTION
Formula SAE (FSAE) is a student collegiate design
series devoted to the design and construction of open-
wheeled race cars. Over 200 schools participate in this
series with competitive events held in the United States
of America, the United Kingdom, Germany, Brazil, Italy,
Japan, and Australia. The event is known for the
enthusiastic participation of student engineers and the
wide range of engineering innovations utilized on the
cars.
The governing rules are limited in technical regulations
except in areas of safety. [1] Although there exist many
others avenues of exploration in engine package design,
a large majority of teams use a naturally-aspirated four-
cylinder engine derived from available motorcycles. A
teams engine choice is limited to any four-stroke with a
displacement less than 610 cc. The number of cylinders
is unlimited. Rules also require a 20-mm restrictor when
94 or 100 Octane fuel is used and a 19-mm restrictor
when instead E-85 fuel is used.
With four-cylinder Formula SAE engines, the shape,
layout, size, and packaging of intake manifolds vary.
Each team uses different runner lengths, plenum
volumes, restrictor designs, etc for their intake manifold
along with different exhaust configurations, which do not
allow for direct comparisons. While the concepts of
intake manifolds vary, the restrictor performance is an
important consideration for any intake design.
Each diffuser angle and intake system was simulated
using industry standard engine simulation software
provided by Ricardo. Ricardo WAVE was first used to
model the intake manifold in a full engine model. Several
of these concepts were then run using a WAVE-VECTIS
coupled model. The WAVE simulations modeled a
600cc 4-cylinder Yamaha YZF-R6 engine across the
entire RPM range, while the coupled models were run at
one or a few RPM points, due to the much higher
computational cost. The coupled model simulates the
intake manifold using the Ricardo VECTIS
Computational Fluid Dynamics (CFD) code coupled to
the rest of engine model simulated with WAVE. Thus, in
the coupled model, the intake manifold is represented in
3D, instead of 1D.
2
The primary investigation focuses on the effect of
different diffuser angles on performance using the
WAVE-VECTIS coupled engine simulation. The diffusers
simulated utilized 7, 5.5, 4, and 3 half-angles. The 3
half-angle restrictor outperformed the 7, 5.5, and 4
half-angle restrictors by improving the volumetric
efficiency of the engine. Changing the diffuser half-angle
by 4 changed the volumetric efficiency (VE) of the
engine investigated at high RPM by 4.0% for the same
plenum configuration.
The diffuser performance is compared against a
theoretical choked flow maximum volumetric efficiency
to understand the ultimate potential of the intake. The
authors demonstrate that the engine may approach and
even exceed the theoretical maximum volumetric
efficiency depending on the tuning of the intake. The
flow fields inside the diffuser (for half-angles of 7, 5.5,
4, and 3) are examined using VECTIS in coupled
simulations with WAVE to gain an improved
understanding of the interactions between the throat,
diffuser, and pressure pulsations. Items such as Mach
number, turbulent kinetic energy, and total pressure are
presented and examined through cycle time averaged
and crank angle resolved plots.
Coupled results indicate that the restrictor is not choked
100% of the time, even at high engine speeds.
Restrictors with tighter diffuser half-angles exhibited
closer attainment of the 100% choked flow ideal. At
14,000 RPM the amount of the time the throat was
choked varied from 47.9% to 83.3% of the complete
cycle.
Turbulent kinetic energy in the diffuser was examined as
a source of flow losses. Tighter diffuser half-angles had
lower levels of turbulent kinetic energy and smaller loss
of total pressure. A rise in total pressure occurred in the
diffuser and possible explanations were considered.
Explanations examined include Reynoldss Number
effects, numeric error, data analysis area averaging
methods, and pulsed flow effects. Real pulsed-flow
effects and errors introduced by area averaging on non-
uniform flows are the most likely causes of total pressure
rise.
The authors also considered acoustic filtering devices
such as Helmholtz resonators to dampen the pressure
wave fluctuations that reach the restrictor throat. The
impact of the Helmholtz resonators on volumetric
efficiency was encouraging. If a Helmholtz resonator
was placed after the restrictor throat there were gains in
VE across a range of RPM. When an inline Helmholtz
resonator was placed before the restrictor throat, gains
in VE were limited to a very narrow RPM band and VE
decreased below the baseline elsewhere across the
range of RPM.
A flow control concept was also considered upstream of
the restrictor throat. Flow control and separation delay
was attempted by the introduction of swirl vanes. The
swirl vanes had little impact on intake performance, but
perhaps more aggressive vanes could offer a slight
benefit.
BACKGROUND
INTAKE BACKGROUND The Formula SAE event
limits the output of the students engines by imposing the
use of a restrictor on the intake air flow. The design of a
restrictor may be as simple as an orifice plate or as
complex as a converging/diverging duct with choice of
design left up to the students. The restrictor throat
diameter limits the ultimate amount of air flow into the
system, but the designers attempt to minimize any
additional flow losses, which occur mostly downstream
of the throat.
In a previous paper the authors looked at various types
of intake system designs used in the Formula SAE
event. [2] Three types of intake concepts were evaluated
and it was determined that a Conical-Spline Intake
provided the best overall solution for engine
performance and cylinder-to-cylinder volumetric
efficiency balance. Thus the main focus of restrictor
development analysis was done with the Conical-Spline
Intake style. A description of the Conical-Spline Intake
design is provided below.
Conical-Spline Intake Design The conical intake
manifold is characterized by the placement of the runner
inlets in a radial symmetric fashion about the main axis
of the plenum. As the restrictor is inline with the plenum,
all the runner inlets are symmetric to the main flow axis
of the restrictor. All four runners are equidistant from the
restrictor throat which is beneficial to the performance of
the restrictor. This arrangement provides more uniform
pressure pulses arriving at the throat and very even
cylinder-to-cylinder volumetric efficiency. The restrictor is
typically in the center of the row of intake valve ports
common to inline four-cylinder engines. As a
consequence of this runner placement and the use of an
inline four-cylinder engine, the runners have to be
curved to mate to the cylinder head ports. The Conical-
Spline Intake models used in this paper use straight
runners due to previous work and to focus on the
interaction of plenum shape and restrictor geometry.
Previous work showed that very short runners with high
amounts of curvature on a Conical-Spline Intake may
induce additional cylinder-to-cylinder volumetric
efficiency imbalance. [2] The intake runner form used
with this intake was designed to be used with a Yamaha
YZF-R6 engine. In addition, the runner has a slight
taper.
Some conical plenums have a true conical shape and
are merely an extension of the diffuser angle, whereas
others have a spline profile that approaches a conical
shape. Previous work determined that continuing the
diffuser out to very large area ratios by way of making
the intake plenum an extension of the diffuser offered
3
little benefit. The assumption that smoothly continuing
the diffuser into the plenum would limit total pressure
loss appeared to be false as most of the change in total
pressure occurred before an area ratio of 8. Maintaining
the assumption that large diffusers are essential to
restrictor performance may hinder packaging of the
intake. [2] Within this paper the authors refer to a conical
intake with a spline profile as a Conical-Spline Intake.
Figure 1 shows an example of a conical type intake a
constant taper cross section profile of the main plenum
volume. Figure 2 shows an example of a conical type
intake with a spline cross section profile of the main
plenum volume. The basic simulation model (from
WAVE) used in the research is provided in Figure 3.

Figure 1. Conical Intake Design Michigan State
University

Figure 2. Conical Intake Design South Dakota School
of Mines
TUNING EFFECTS OF THE RESTRICTOR Since the
computation expense of 1D/3D coupled simulations is
very high, the authors decided that the problem should
first be examined in 1D to better understand the design
space and the effect of various parameters on
performance. When looking at only a few RPM points
from a 1D/3D solution, it is difficult to ascertain the
effects of diffuser half-angle, length, etc on the
volumetric efficiency curve. The angle of the diffuser
affects engine volumetric efficiency, especially at mid to
high RPM. The diffuser impacts volumetric efficiency
from not only flow and shock losses, but also by
impacting pulse tuning. An increase in volumetric
efficiency between two restrictors, compared at only one
RPM point, could be from reduced flow losses, tuning
effects, or both. While the two effects can never be
totally separated, they must be each examined to gain
an unbiased insight into improving the overall design.
Figure 3. Conical-Spline Intake Basic Simulation
Model
The shape, angle, length, and positioning of the diffuser
all impact the VE curve of an engine. The mere
requirement of physically packaging a restrictor also has
a large impact on the intake manifold shape and layout.
In the case of a restricted engine, the geometry of the
diffuser after the restrictor throat and the geometry of the
plenum are considered important for pressure recovery.
Limiting the change in plenum volume is important as
that will have some effect on the intake system
performance and the overall engine volumetric
efficiency. In general, an increase in volume after the
restrictor throat will often tend to increase volumetric
efficiency high in the RPM band. Limiting the change in
plenum volume is important as that will have some effect
on the intake system performance and the overall
engine volumetric efficiency.
4
DEFINTIONS OF INTAKE MODEL GEOMETRY WITH
VARYING DIFFUSER ANGLE
As the diffuser half-angle is changed, other geometry
parameters of the intake are also changed, making a fair
comparison sometimes difficult. For example, if the
exact same plenum is to be used while varying the
diffuser half-angle, the length of the diffuser will have to
change. In this case the diffuser exit diameter must
match the plenum diameter where the two parts mate
together. However if the diffuser length is held constant
and the diffuser half-angle is varied, the result in this
case is the diffuser exit diameter must change.
With the focus of the simulations being on diffuser half-
angle, it becomes essential to define the limits of the
geometries researched. The diffuser length is the
distance from the throat to the diffuser exit. The diffuser
exit is simply the end of the diffuser and where the
plenum starts. The volume of the diffuser is the volume
contained in the diffuser from the throat plane to the
diffuser exit plane.
To define the plenum volume, arbitrary surface
boundaries were defined. The diffuser exit plane defines
the upper boundary. The plane which defines the bottom
of the plenum, or the very beginning of the runners,
defines the lower boundary. Thus, everything after the
exit plane of the diffuser and before the entry to the
runners is considered as the plenum volume. The
plenum length is the distance from the diffuser exit plane
to the bottom of the plenum.
The combinations of diffuser and plenum geometry are
also defined. The total length is the length of the diffuser
plus the plenum length. The total volume is the volume
of the plenum plus the volume of the diffuser. Thus the
total volume is an indicator of all of the intake volume
downstream of the throat, not including the runners.
The restrictors in the simulations utilized 7, 5.5, 4, and
3 diffuser half-angles. These diffuser half-angles were
investigated in WAVE using the Conical-Spline Intake
and a variant of the Conical-Spline Intake. Various
geometry configurations were tested, including changing
diffuser half-angle while using the same plenum. The
restrictors used in the WAVE-VECTIS simulations
utilized 7, 5.5, 4, and 3 diffuser half-angle, all with
the exact same plenum. The various restrictors were
then examined by looking at items such as volumetric
efficiency, Mach number, turbulent kinetic energy, and
total pressure along the diffuser.
WAVE RESULTS OF DIFFERENT RESTRICTOR
DIFFUSER ANGLES
IMPACT OF DIFFUSER ANGLE ON VOLUMETRIC
EFFICIENCY USING WAVE As this is a 1D analysis,
there are no 3D flow loss effects in the diffuser. Tuning
effects and shock losses will still be accounted for
however. In this regard, the absence of separation
losses in this analysis, actually allowed the authors to
look more at the pure tuning effects of the diffuser, and
its interaction with the plenum.
Changes in both length and volume affect the overall
tuning of the intake. The authors tried to understand
these aspects on the intake so the effects of the various
diffuser angles could be separated out from just the pure
changes in length and volume. Two cases are examined
using the Conical-Spline Intake with varying restrictor
half-angles. A third case is examined using a similarly
shaped intake manifold.
Case 1: Intake With Constant Plenum Volume and
Height In Case 1, the exact same plenum was used for
all the diffuser half-angles simulated. As the same
plenum was used, and all of the diffusers must mate to
the same plenum, this fixed the diffuser exit diameter
dimension. One geometric effect of different diffuser
half-angles is as the diffuser angle is increased from 3
to 7, the length of the diffuser decreases. The plenum
volume and plenum length are held constant. Both the
total length and total volume after the throat changes for
each restrictor angle, but the change is caused by the
use of different diffuser angles. The relative dimensions
are noted in Table 1 below.
Diffuser Half-Angle 3 4 5.5 7
Diffuser Exit Diameter (mm) 72.90 72.90 72.90 72.90
Diffuser Length (mm) 504.7 378.2 274.7 215.4
Diffuser Volume (liters) 0.95 0.71 0.52 0.40
Plenum Volume (liters) 2.26 2.26 2.26 2.26
Plenum Length (mm) 160.5 160.5 160.5 160.5
Total Length from Throat to
Plenum Bottom (mm)
665.2 538.7 435.2 375.9
Total Volume After Throat
(liters)
3.21 2.97 2.78 2.66
Table 1. Dimensions of Intake and Restrictors for Case 1
Looking at Table 1, while the plenum volume remains
constant at 2.26 liters, the total volume increases with
decreasing half-angle. The total volume increases from
2.66 liters using a 7 half-angle to 3.21 liters using a 3
half-angle. Thus from the 7 to the 3 restrictor, the total
volume after the throat increases by 20.7%. Also, while
the plenum length has not changed, the total length
(distance from throat to plenum bottom) is changing for
each half-angle. The total length increases from 375.9
mm for the 7 restrictor to 665.2 mm for the 4 restrictor.
This is a 77% increase in total length. Thus the total
length and post-throat volume have all changed as the
half-angle is varied.
Figure 4 shows the change in VE as the diffuser half-
angle is changed, using the same plenum. Looking at
Figure 4, there are significant shifts in the VE curve
across the rev range of the engine. As the RPM goes
up, the VE curves tend to come together.
5
The 5.5 restrictor has more volume than the 7, yet the
5.5 shows a much lower VE from 11,000 to 14,000
RPM. The 4 restrictor also shows a lower VE than the
7 restrictor from 12,250 to 13,000 RPM. This indicates
there is more to the determination of a VE curve than
plenum volume or the total volume after the throat.
Figure 4. Case 1: Volumetric Efficiency for Various
Diffuser Half-Angles with Constant Exit Diameter
The fact that the length is changing with different half-
angles is not trivial in regard to the VE curve. Figure 5
shows multiple order plots for the Conical-Spline Intake
with a 7 half-angle diffuser at 14,000 RPM. 2
nd
and 4
th

orders are most dominant. Looking at the order plots,
one can see a standing pressure wave is set up within
the intake and diffuser along the main length of the
plenum. The pressure anti-node is located just
downstream of the throat. Any pressure wave is partially
reflected back by a reduction in area. The diffuser is a
reduction in area when seen by the direction of wave
propagation from the runners. The wave source can be
thought of as the runners and part of the energy is
reflected back to the source by the diffuser. As a side
note, it can be seen that all the odd orders (i.e. half
engine orders) exhibit cross modes. All the odd orders
were much lower amplitude than the even orders.
Case 2: Intake With Constant Diffuser Length and
Differing Diffuser Exit Diameter In Case 2, the length
of the diffuser was not allowed to change, and was fixed
to 216.4 mm from the throat to the diffuser exit. The
plenum length was also fixed to the same length as in
Case 1, at 160.5 mm. In Case 2, the diffuser exit
diameter was determined by the diffuser angle and the
aforementioned predetermined diffuser length of 216.5
mm. As the exit diameter changed for each case, the
plenum was adjusted slightly to keep the plenum volume
constant at 2.25 to 2.26 liters for each case. While the
distribution of volume (i.e. the cross section) along the
length of the plenum will differ slightly for each diffuser
angle case, the plenums are very similar. The base
diameter of the plenum was also not changed. Table 2
shows the relative dimensions for Case 2. From Table 2
notice that the total volume is decreasing with
decreasing half-angle. Thus Case 2 effectively has the
opposite effect of Case 1 with regard to total volume.

Figure 5. Pressure Amplitude for Orders 1 8, Case 1
Conical-Spline Intake with 7 Diffuser at 14,000 RPM

Diffuser Half-Angle 3 4 5.5 7
Diffuser Exit Diameter (mm) 42.578 50.126 61.485 72.898
Diffuser Length (mm) 215.4 215.4 215.4 215.4
Diffuser Volume (liters) 0.17 0.22 0.31 0.40
Plenum Volume (liters) 2.25 2.25 2.26 2.26
Plenum Length (mm) 160.5 160.5 160.5 160.5
Total Length from Throat to
Plenum Bottom (mm)
375.9 375.9 375.9 375.9
Total Volume After Throat
(liters)
2.423 2.474 2.564 2.663
Table 2. Dimensions of Intake and Restrictors for Case 2
6
In Figure 6 there is a general trend of VE increasing in
the range of 11,000 to 13,000 RPM as the restrictor half-
angle is decreased. The 3 half-angle restrictor also
shows the largest increase even though it has 9% less
total volume than the 7 half-angle restrictor.
Figure 6. Case 2: VE of Conical-Spline Intake with
Different Diffuser Angles and Constant Total Intake
Height
In contrast to Case 1, the VE curves in Case 2 generally
have the same shape and appearance and have small
differences in VE. When volume changes are made to
the diffuser, the need exists to evaluate if VE changes
are due solely from the diffuser change or other
concurrent plenum changes. If changes in plenum
volume are considered in restrictor performance, one
must also evaluate the impact of length or geometry
changes to obtain that volume increase. Case 1 and
Case 2 illustrate this point. Furthermore Case 1 and
Case 2 show the difficulty in comparing multiple intakes
at only a few rpm points if some concept of the entire VE
curve is not known.
Case 3: Simplified Intake With Differing Cylindrical
Plenum Volumes To further investigate the
aforementioned parameters such as plenum length,
diffuser length, and plenum volume, a cylinder style
intake manifold was used instead of the Conical-Spline
Intake manifold. The reason for this is that a cylinder has
a constant distribution of cross-section along its length,
and is much more readily and quickly modified in
WAVEs meshing program, WaveBuild3D. Figure 7
shows the general shape of the cylinder type manifolds.
The cylinder type manifold also exhibits very similar VE
curves and very low cylinder-to-cylinder VE imbalance,
like the Conical-Spline Intake. All the following
simulations (Figures 8 through 10) use a restrictor with a
7 half-angle diffuser. Thus in the following plots only the
changes of plenum volume, length, and diameter impact
the VE curve. The main goal here is to determine how
much of a role is played purely by the plenum volume,
versus the distribution of that volume. All runs used an
automatically meshed intake to make sure pressure
wave modes were well represented.
Figure 7. Cylindrical Intake
Figure 8 shows the effect of increasing volume while
holding plenum length constant at 160.5 mm. Plenum
volume is increased by merely increasing the cylinder
diameter. The minimum volume was limited by the layout
of the runners, which did not change from the Conical-
Spline Intake. Looking at Figure 8, one can see that VE
does increase with increasing plenum volume, but only
slightly. By more than doubling the intake volume, the
VE has increased by less than 1% over most of the rev
range. Also note the striking similarity to Figure 6. Both
exhibit changes in VE around 11,500 RPM.
Figure 8. Cylinder Intakes of Differing Volume at
Constant Plenum and Diffuser Length
Using the cylinder type manifold again, Figure 9 shows
the effect of holding the plenum volume constant. Here
the plenum length is allowed to change, from 78.3 mm,
7
to 160.5 mm, to 187.15 mm. The same 7 restrictor is
used again for all of the data in Figure 9. With slightly
more than a doubling of the plenum length, the VE peak
has shifted from 11,500 RPM down to 11,250 RPM, and
peak VE has increased by 0.9%. Also, Figure 9 shows
larger RPM shifts in the VE peaks than Figure 8.
Figure 9. Constant Volume (3.9 l) Cylinder Intake with
Variable Plenum Length
Figure 10 examines the same effect as Figure 9, but this
time with an 8.0 liter plenum volume. As the plenum
volume was larger, there was a larger range in which to
shift the plenum height, while still retaining the 8.0 liter
volume. By doubling the plenum height, the VE has
increased by 1.2%, and has been shifted down 750
RPM.
Figure 10. Constant Volume (8.0 l) Cylinder Intake
with Variable Plenum Length
The lesson learned from the study of Cases 1 through 3
is the effect of geometry changes to diffuser and
plenum, result in changes to performance. These
changes in performance are not simply additive, but in
fact tend to be highly non-linear. In Case 1, the variation
of overall intake length, volume and diffuser angle
resulted in a 6% peak variation in VE at mid engine
speeds. Furthermore at the highest rpm shown (15,000
RPM), the intake with the highest total volume did not
show the highest VE. Case 2 only looked at changes in
diffuser angle, where the peak variation in VE was under
2%, but tighter diffuser half-angles conclusively had
higher VE. In Case 3, different plenum volumes and
lengths were examined with the same restrictor and in
one configuration the peak variation in VE was only
1.2%. Contrasting the VE curves of Case 1 with Case 2
and 3 shows the complexity of trying to optimize the
overall intake geometry to maximize VE, and how the
restrictor plays a role in the overall VE curve.
Transient runs were also conducted in WAVE using
rapid acceleration parameters of RPM and throttle to
determine if there would be any negative impact to
throttle response or VE due to the change in total
plenum volume and/or diffuser dimensions. WAVE
indicated almost the same exact trend in VE curves as
with the steady state run for all cases. Above 13,500
RPM during a transient run the cases with greater total
volume did show an improvement in VE of 1% to 2%
over the steady state simulation. A transient run can be
thought of as running a real engine on a roller/inertial
dynamometer. All other WAVE runs were simulated
steady state as if the engine was on a dynamometer that
can hold a constant RPM. Steady state in this case does
not mean fixed boundary conditions as used in a steady
state CFD simulation.
WAVE-VECTIS BACKGROUND AND RESULTS
OF DIFFERENT RESTRICTOR DIFFUSER
ANGLES
The use of 1D WAVE simulations allowed the authors to
look at the wave pulsation effects throughout the intake
system, but the flow within the intake is inherently a 3D
flow field and requires the use of CFD. A WAVE-VECTIS
coupled model to get a more in-depth look at the flow
field inside the diffuser on the Conical-Spline Intake. A
coupled approach was used to avoid inaccuracies in a
steady-state flow approach as learned in previous work.
[4]
WAVE-VECTIS Boundary Conditions The boundary
conditions in the WAVE model were fixed to 300 K and
100 kPa at the inlet boundary and floating at the exhaust
boundary. VECTIS boundary conditions were
determined by the WAVE calculation at the inlet and
exits for the intake manifold. VECTIS uses a prescribed
mass flow forcing at the boundaries as determined by
WAVE. The five species present in parts of the flow are
air, fuel vapor, burned air, burned fuel, and liquid fuel.
Continuity is preserved across WAVE-VECTIS
boundaries for all five species. The walls of the concept
intakes modeled in VECTIS had a specified temperature
of 300 K and a perfectly smooth wall surface roughness.
8
In an effort to reduce the effect of forcing the boundary
conditions on the results of the 3D CFD simulation, the
intake manifold included additional features. A box was
added, with dimensions of 152.4 mm per side, placed
near the inlet of the intake where the WAVE and
VECTIS codes meet. See Figure 11. In addition, a
significant portion of the overall runner length was
modeled with VECTIS rather than WAVE. Both these
features had the effect of keeping the boundaries far
away from the complex 3D flows and providing a uniform
1D flow at the junction between the codes.
Figure 11. Standard Inlet Box Attached to Intake to Move
the WAVE-VECTIS Boundary Away From the Main Area
of Interest
Fuel Injector Placement Formula SAE manifolds
typically have the fuel injectors mounted on the intake
runners. The VECTIS CFD model for the intakes in this
paper included the entire runner length of the intake
manifold, up to the cylinder head interface. In order to
keep computational cost and complexity to a minimum,
the fuel injector placement was not modeled with the
injector on the intake manifold runner. Having the
injector included in the VECTIS model would mean that
the fuel spray would have to be included in the 3D CFD
calculations, which would increase the complexity of the
VECTIS model. Instead, the fuel injectors were placed in
a WAVE duct that represents the immediate entry to the
cylinder head intake ports. The effect of this is that the
fuel injectors are placed within the WAVE model
approximately 20 mm downstream from the injectors
normal position, which has minimal changes in the
model output. In addition, this change eliminated the
need to model the injector mounting boss and
associated injector hole within the runner portion of the
CFD model.
Elimination of Throttle Plate It was decided to neglect
the modeling of the throttle plate common to
conventional throttle bodies upstream of the restrictor in
order to reduce mesh size and computation time. This
was felt to be reasonable in that all four restrictor angles
would still be modeled on equal footing. It has been
found by other researchers that complete modeling of
the butterfly valve had little effect on overall performance
results at WOT; however this was with a non-restricted
engine. [3]
Computers Used and Typical Runtime Simulations
were carried out on an IBM SP and an IBM Regatta
running multiple processor nodes at the Minnesota
Supercomputing Institute. Simulations took
approximately 2500-3400 CPU hours on the IBM
Regatta, for the higher RPM data points. Higher RPM
runs required more cycles to converge. The number of
CPU hours required depended on the RPM of the
simulation and the capabilities of the machine. Models
contained up to 600,000 cells, which correlates to about
a 3.5 mm on a side mesh cell size and a 1.75 mm on a
side cell size in the restrictor throat and diffuser area. It
was found that the VECTIS code offered fairly linear
scalability with an increasing number of processors.
Convergence of volumetric efficiency was held to less
than 0.5% per each individual cylinder. Convergence
was obtained in 16 to 23 WAVE-VECTIS coupled engine
cycles.
IMPACT OF DIFFUSER ANGLE ON VOLUMETRIC
EFFICIENCY USING WAVE-VECTIS A WAVE-
VECTIS analysis captures more realistic flow and flow
losses than the simpler 1D WAVE-only analysis. These
additional flow losses result in an additional loss of
volumetric efficiency. The intake and diffuser geometry
of Case 1 form the basis for the WAVE-VECTIS
simulation. To help in comparing the WAVE and WAVE-
VECTIS results, both results are compared to a
theoretical maximum volumetric efficiency that the
engine may obtain for a restricted intake system.
Finding the maximum theoretical volumetric efficiency
requires that the maximum flow through the throat be
estimated. The maximum flow through an adiabatic and
isentropic 1D orifice, once choked, is given by Equation
1. [4] Once choked flow is achieved, the mass flow
through the orifice is limited, except by increasing the
upstream stagnation density or pressure. The flow
demand of the engine is given by Equation 2. The
theoretical maximum volumetric efficiency is then a
comparison of what the restrictor can deliver versus the
demand of the engine, which is given by Equation 3.
( ) ( )
0
0 0
1 1
*
1
2


p A
Q
throat
choked
+
|
|
.
|

\
|
+
=
&
(1)
Second Per Cycles nt Displaceme Q
nom
* *
2
1
.
=
&
(2)
.
max
. .
nom
choked
Q
Q
E V
&
&
= (3)
Inlet Box
Inlet Face
9
The theoretical maximum volumetric efficiency is
compared to the WAVE and WAVE-VECTIS results in
Figure 12. Some data points closely overlap each other
in Figure 12, thus Table 3 shows the actual volumetric
efficiencies.
0.85
0.90
0.95
1.00
1.05
1.10
1.15
1.20
1.25
9000 10000 11000 12000 13000 14000 15000
RPM
V
O
L
U
M
E
T
R
I
C

E
F
F
I
C
I
E
N
C
Y

Theoretical Maximum
7 Degree - WAVE
7 Degree - WAVE-VECTIS
5.5 Degree - WAVE
5.5 Degree - WAVE-VECTIS
4 Degree - WAVE
4 Degree - WAVE-VECTIS
3 Degree - WAVE
3 Degree - WAVE-VECTIS
Figure 12. Theoretical Maximum Volumetric Efficiency
Compared to the WAVE and WAVE-VECTIS
Simulations
Beyond approximately 12,000 RPM, the WAVE
simulations display a VE that is above the theoretical
maximum VE. However, some of the WAVE-VECTIS
data points also exceed the theoretical maximum VE, as
shown in Figure 12 and Table 3. If the diffuser and
intake system are designed well, then the system should
approach the maximum theoretical volumetric efficiency
at high RPM. Looking at Table 3, the 5.5 half-angle
diffuser comes within 0.1% of matching the theoretical
maximum VE, whereas the 4 half-angle diffuser
exceeds the theoretical maximum VE by 0.9%. The 3
half-angle diffuser exceeds the theoretical maximum VE
by 1.3%.
Theoretical
Maximum
Wave-
Vectis
Wave
Difference
(Wave - W-V)
7 10 126.5% 97.2% 103.7% 6.5%
4 10 126.5% 101.3% 105.3% 4.0%
7 11.5 110.0% 100.5% 107.9% 7.4%
7 14 90.4% 87.8% 93.7% 5.9%
5.5 14 90.4% 90.3% 93.2% 2.9%
4 14 90.4% 91.3% 92.9% 1.6%
3 14 90.4% 91.7% 94.3% 2.6%
Half
Angle
RPM
x1000
VOLUMETRIC EFFICIENCY
Table 3. Comparison of WAVE and WAVE-VECTIS
Simulations to Theoretical Maximum Volumetric
Efficiency
While the maximum theoretical VE gives a good
indicator of the performance achievable for a particular
engine, especially as RPM increases, it is not the final
word on the maximum VE achievable. Equation 3 makes
an implicit assumption wherein the unsteady, cycle
dependent flow through the valves is deemed to be
equal to the steady state flow through the restrictor. The
flow through the valves is highly unsteady, with pulse
tuning of the runners and intake plenum also occurring.
The pressure during intake valve closure, due to cam
timing and pressure wave tuning is of prime importance
in dictating VE. Thus looking purely at the steady flow
through the throat is too simplistic to analyze the
maximum obtainable VE.
The implicit assumption is made by Equations 1 and 3
that the flow is choked 100% of the time, which does not
occur in engine operation even at 14,000 RPM
according to simulation results. The 7 half-angle
restrictor at 10,000 RPM does not achieve choked flow
(as will be shown later), yet there is still a large disparity
in volumetric efficiency between the WAVE and WAVE-
VECTIS solution. However both the WAVE and WAVE-
VECTIS solutions for the 7 half-angle restrictor at
10,000 RPM are still well below the theoretical maximum
VE. Thus at 10,000 RPM there are more gains to be
obtained, likely through flow loss reductions and tuning,
even though the restrictor does not reach Mach 1 during
the cycle.
The WAVE-VECTIS simulations all display a VE lower
than WAVE. The result that WAVE predicts a higher VE
in relation to WAVE-VECTIS is expected, as WAVE
simply can not predict 3D flow losses. Looking at Table
3, the very right hand column shows the difference in VE
between WAVE and WAVE-VECTIS. The VE difference
helps to judge how close each restrictor is to realizing its
individual VE potential at each RPM point. Comparing
the various restrictors at 14,000 RPM, it can be noted
that the 7 restrictor has more VE difference than any of
the other restrictors. At 10,000 RPM the VE difference
for the 7 restrictor is also greater than the 4 restrictor.
This at least implies 3D flow losses are decreasing as
restrictor half-angle is decreasing.
The following three sections go into more detail
regarding Mach number, turbulent kinetic energy, and
total pressure, in the diffuser section.
MACH NUMBER IN THE DIFFUSER The WAVE-
VECTIS results in this section and the following two
sections are plotted as a function of a normalized area
ratio. The area ratio is given by Equation 4 where
x
A
is
the cross sectional area of the diffuser at some distance
downstream of the throat and
*
A
is the area at the
throat. For reference, the restrictor throat has a
normalized are ratio of 1, and the restrictor exit has a
normalized area ratio of 13.29.
*
A
A
Area Normalized
x
= (4)
To plot the data from the VECTIS portion of the
simulation, multiple planes were set up perpendicular to
the restrictor flow axis. Data was extracted at each plane
10
in 5 crank-angle degree (CAD) increments. Data from
each plane was extracted as a planar area average of
the variable of interest, whether the variable was total
pressure, Mach number, or turbulent kinetic energy. For
time averaged plots, data was extracted in the same
manner, however one cycle was time averaged over the
last cycle of each run. Thus the time averaged runs are
an average of the whole 720 CAD of the engine cycle.
The cycle time averaged Mach number along the
diffuser is shown below in Figure 13 for each diffuser
half-angle. Looking at Figure 13, it can be seen that all
the diffusers generally achieve their time averaged peak
Mach number at a normalized area ratio of about 1.15.
Planar Mean Mach Number at 14,000 RPM
(Time Averaged)
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1.0
1.1
1.2
1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0 5.5 6.0
Normalized Area
M
a
c
h

[
#
]
7 5.5 4 3
Figure 13. Cycle Averaged Mach Number for the 7,
5.5, 4, and 3 Restrictors as a Function of Normalized
Area Ratio of the Diffuser at 14,000 RPM
Figures 14 to 17 show pictures for the cycle time
averaged Mach number for the various half-angles along
the center plane of the restrictor. Figures 14 to 17 have
the plot scale adjusted to show the time averaged peak
Mach number. Figures 14 to 17 show each restrictor to
have highly separated flow. However the 7 half-angle
restrictor shows a less axially centered flow. Also, the
5.5, 4, and 3 half-angle restrictors all exhibit similar
time-averaged peak Mach numbers.
Figure 14. Cycle Time-Averaged Mach number for 7
Restrictor at 14,000 RPM
Figure 15. Cycle Time-Averaged Mach Number for 5.5
Restrictor at 14,000 RPM
Figure 16. Cycle Time-Averaged Mach Number for 4
Restrictor at 14,000 RPM
Figure 17. Cycle Time-Averaged Mach Number for 3
Half-Angle Restrictor at 14,000 RPM.
Mach Number Flow Uniformity in the Diffuser Figure
18 shows the flow uniformity index [5] of Mach number
for all the restrictors at 14,000 RPM. A flow uniformity
index of 1 is equal distribution of a scalar variable
among all cells in a plane. The flow uniformity index is
defined by Equations 5 and 6.
11
mean
mabs

2
1 = (5)
Where mabs is defined as the mean absolute deviation
from the mean as defined below.


=
i
mean i i
A
A
mabs

(6)
i
A is defined as the area of the mesh cell i , cut by an
arbitrary plane of interest.
i
is a scalar value in the
mesh cell i and
mean
is the mean value of the scalar
over the entire plane of interest.
Above an area ratio of 2 to 3, there is a general trend
that the lower half-angle restrictors have better flow
uniformity. Flow uniformity also is almost the same
between all cases, until the curves start diverging
around an area ratio of 1.5 to 1.6. Also the 7 half-angle
restrictor is the only restrictor that does not show
improving flow uniformity in the downstream section of
the diffuser. All other half-angles show improving flow
uniformity as the flow moves downstream.
Flow Uniformity of Mach Number at 14,000 RPM
(Time Averaged)
0.70
0.75
0.80
0.85
0.90
0.95
1.00
1 2 3 4 5 6 7 8 9 10 11 12
Normalized Area
F
l
o
w

U
n
i
f
o
r
m
i
t
y

I
n
d
e
x

7 5.5 4 3
Figure 18. Cycle Averaged Flow Uniformity Index for
Mach Number for the 7, 5.5, 4, and 3 Restrictors as
a Function of Normalized Area Ratio of the Diffuser at
14,000 RPM
Figure 19 shows the flow uniformity index of Mach
number for the 7 and 4 half-angle restrictors at 10,000
RPM. Notice Figure 19 has a different scale than Figure
18. For the 7 half-angle restrictor the flow uniformity
index of Mach number is lower at 10,000 RPM than at
14,000 RPM above an area ratio of 2.5. The 4 half-
angle flow uniformity index has generally not changed
much from 14,000 to 10,000 RPM. However, a slight
increase in flow uniformity index is noticeable below an
area ratio of 4.
Flow Uniformity of Mach Number at 10,000 RPM
(Time Averaged)
0.60
0.65
0.70
0.75
0.80
0.85
0.90
0.95
1.00
1 2 3 4 5 6 7 8 9 10 11 12 13
Normalized Area
F
l
o
w

U
n
i
f
o
r
m
i
t
y

I
n
d
e
x
_
7 4
Figure 19. Cycle Averaged Flow Uniformity Index for
Mach Number for the 7 and 4 Restrictors as a
Function of Normalized Area Ratio of the Diffuser at
10,000 RPM
While the time averaged plot and pictures provide
information about the various restrictors, they still leave
some information to be desired, as the crank angle
resolved details are averaged out. Figures 20 to 25 plot
the Mach number over one full engine cycle at various
normalized area ratios. Looking at the plots, the reader
can quickly see that the flow is not choked 100% of the
time.
In any engine, pressure waves are traveling through the
intake manifold generated by the unsteady flow through
the intake valves. In a restricted engine, even at high
RPM, these pressure waves reach the throat. The
velocity in the throat at any time during the cycle is
largely a function of the instantaneous pressure ratio
across the throat. It is often stated that once the velocity
in the throat reaches Mach 1, that the pressure waves
can not travel upstream of the throat, except through the
subsonic boundary layer. While it is true that the flow is
choked at Mach 1, it must be remembered that the
pressure ratio is driving this situation, not the other way
around.
As the RPM increases, the Mach number becomes
supersonic due to an intake suction pulse reaching the
throat. The percentage of the cycle that is choked also
increases as RPM increases, which makes sense. Table
4 shows the percentage of the cycle that is choked for
each diffuser half-angle and RPM. Note that at 14,000
RPM the 3 restrictor is choked 83% of the cycle time,
whereas the 7 restrictor is choked only about 48% of
the cycle time, keeping in mind that the 3 restrictor
outperformed the 7 restrictor in VE.

12
Figure 20. Mach Number at 10,000 RPM Over One
Engine Cycle for the Conical-Spline Intake with 7
Restrictor
Figure 21. Mach Number at 10,000 RPM Over One
Engine Cycle for the Conical-Spline Intake with 4
Restrictor
Figure 22. Mach Number at 14,000 RPM Over One
Engine Cycle for the Conical-Spline Intake with 7
Restrictor

Figure 23. Mach Number at 14,000 RPM Over One
Engine Cycle for the Conical-Spline Intake with 5.5
Restrictor
Figure 24. Mach Number at 14,000 RPM Over One
Engine Cycle for the Conical-Spline Intake with 4
Restrictor
Figure 25. Mach Number at 14,000 RPM Over One
Engine Cycle for the Conical-Spline Intake 3 Restrictor
13
Notice that as the diffuser half-angle decreases, the
peak Mach number increases, but the peak Mach
number achieved occurs for less and less time during
the cycle. While increasing the Mach number should
increase shock losses, the shock losses are a
combination of both the Mach number at which the
shocks occur, plus the percentage of time at that Mach
number for which they occur.
Restrictor
Half Angle
RPM
% of Cycle
Choked
7 10,000 0%
4 10,000 15.3%
7 11,500 0%
7 14,000 47.9%
5.5 14,000 68.0%
4 14,000 80.0%
3 14,000 83.3%
Table 4. Percent of Cycle Choked for Various Restrictor
Half-Angles and RPM.
To gain a further insight into the crank angle resolved
Mach numbers, Figures 26 to 29 show the Mach number
at various points in the engine cycle for the 7 and 4
half-angle restrictors. The maximum Mach numbers
achieved for the 7 and 4 half-angle restrictors are
plotted in Figures 26 and 27, respectively. The
preceding cycle minimum Mach number at the throat
was then determined and plotted in Figures 28 and 29.
Figure 26. Mach Number for the 7 Restrictor Along
Center Plane at 455 Crankangle Degrees
The 7 half-angle restrictor displays swirling flow within
the diffuser for both the maximum and minimum Mach
numbers during the cycle. During the minimum Mach
number, the 7 half-angle restrictor also displays
separated flow quickly after the throat. While the flow in
the 4 half-angle restrictor also displays separation, it is
not as pronounced as in the 7 half-angle restrictor.
However, the 4 half-angle restrictor does not display the
same high degree of swirling of the flow in the
downstream section of the diffuser as compared to the
7 half-angle restrictor.
Figure 27. Mach Number for the 4 Restrictor Along
Center Plane at 510 Crankangle Degrees
Figure 28. Mach Number for the 7 Restrictor Along
Center Plane at 360 Crankangle Degrees
Figure 29. Mach Number for the 4 Restrictor Along
Center Plane at 425 Crankangle Degrees
After the peak Mach number is achieved, the Mach
number decreases rapidly. Figures 30 and 31 display a
time in the cycle where the Mach number is rapidly
decreasing. This is an area of rapidly changing pressure
14
ratio. The point at which the Mach number is decreasing
rapidly is of interest as this is when the peak turbulent
kinetic energy is occurring during the cycle. Figures 30
and 31 are referenced in the following section, which
deals with turbulent kinetic energy within the restrictor.
Figure 30. Mach Number for the 7 Restrictor Along
Center Plane at 515 Crankangle Degrees
Figure 31. Mach Number for the 4 restrictor Along
Center Plane at 540 Crankangle Degrees
TURBULENT KINETIC ENERGY IN THE DIFFUSER
Investigating turbulent kinetic energy (TKE) over the
engine cycle allows some further insight into turbulent
losses and possible areas for further improvement. The
generation of turbulence reduces the flow-work available
and can reduce total pressure available in the intake
system.
Figure 32 displays the time averaged turbulent kinetic
energy for the various restrictors. The 7 and 5.5
diffusers both show a larger amount of turbulent kinetic
energy while the 3 has the lowest turbulent kinetic
energy. Figure 32 also shows a basic trend that after an
area ratio of 3, the time averaged TKE value goes down
as diffuser half-angle is decreased from 7 down to 3.
At area ratios lower than 3, there is a similar shape to
the curves, but it is difficult to determine a trend based
on diffuser half-angle. All the restrictors exhibit virtually
identical TKE values below an area ratio of 1.5.
Planar Mean Turbulent Kinetic Energy at 14,000 RPM
(Time Averaged)
0
200
400
600
800
1000
1 2 3 4 5 6 7 8 9 10 11 12
Normalized Area
T
u
r
b
u
l
e
n
t

K
i
n
e
t
i
c

E
n
e
r
g
y

[
m
2
/
s
2
]
7 5.5 4 3
Figure 32. Cycle Time Averaged Turbulence Kinetic
Energy for the 7, 5.5, and 4 Restrictors at 14,000
RPM
Figure 33 shows the Time Averaged TKE values for the
7 and 4 half-angle restrictors at 10,000 RPM. As the
flow through the throat is subsonic for the 7 half-angle
restrictor, TKE values at the throat and immediately after
are the relative low points. In contrast to 14,000 RPM,
TKE values for the 7 half-angle restrictor at 10,000
RPM are the highest in the farther downstream section
of the diffuser. As the 4 half-angle restrictor is achieving
supersonic flow during a portion of the cycle, the TKE
curve tends to look more like that of the high RPM data
points (as shown in Figure 32). The time averaged peak
TKE value moves upstream (i.e. lower area ratio) with
increasing RPM or increased flow velocity.
Planar Mean Turbulent Kinetic Energy at 10,000 RPM
(Time Averaged)
20
40
60
80
100
120
140
1 2 3 4 5 6 7 8 9 10 11 12 13
Normalized Area
T
u
r
b
u
l
e
n
t

K
i
n
e
t
i
c

E
n
e
r
g
y

[
m
2
/
s
2
]
7 4
Figure 33. Cycle Time Averaged Turbulence Kinetic
Energy for the 7 and 4 Restrictors at 10,000 RPM
TKE over the engine cycle at 14,000 RPM is shown in
Figures 34 to Figure 37. Each curve denotes the TKE at
a particular area ratio within the diffuser. TKE for the 7
restrictor barely peaks over 1800 m
2
/s
2
, while the 5.5
and 4 restrictors both have TKE peaks of around 2500
m
2
/s
2
, while the 3 restrictor has TKE peaks all around
1600 m
2
/s
2
. The 4 has the highest TKE peaks, but it
15
also spends a smaller portion of the cycle at its peak
TKE numbers, as compared to the 7 and 5.5 restrictor.
As the restrictor angle decreases, the flow structure
generally becomes more repeatable. This effect can be
appreciated by the more structured and repeatable cycle
of TKE values seen as the restrictor angle decreases.
The 3 and 4 restrictor have a very repeatable pattern
of TKE during the engine cycle, at area ratio, as where
the 7 restrictor is much less cyclically repeatable. The
5.5 restrictor does have some cyclic variation, but is
closer in this regard to the 4 and 3 restrictors than the
7 restrictor. The uniformly periodic variation of TKE is
related to the variation in Mach number of the cycle as
seen in the previous section. The 7 restrictor showed
the least repeatable pattern of Mach number variation
over the cycle and this is reflected in the TKE plot. Also
notice the TKE values at the throat are virtually identical.
However, after an area ratio of 1.25, the TKE values
differ widely.
Figure 34. Turbulent Kinetic Energy for the 7 Restrictor
Along the Diffuser at 14,000 RPM
Figure 35. Turbulent Kinetic Energy for the 5.5
Restrictor Along the Diffuser at 14,000 RPM
Figure 36. Turbulent Kinetic Energy for the 4 Restrictor
Along the Diffuser at 14,000 RPM
Figure 37. Turbulent Kinetic Energy for the 3 Restrictor
Along the Diffuser at 14,000 RPM
Figures 38 to 45 graphically show the turbulent kinetic
energy through the center plane of the restrictor at
various crank angles in the cycle. Notice that the scale
maximums are not the same for all the pictures.
The CAD points for Figures 38 to 41 were selected to
show a TKE peak during the engine cycle for each
respective restrictor. During this time in the engine cycle,
the Mach number is rapidly decreasing. The TKE peak
occurs at 515 CAD for the 7 restrictor and 540 CAD for
the 4 restrictor. Mach plots at these exact CAD points
were shown earlier in Figures 30 and 31.
Figures 42 to 45 show TKE at a point 90 CAD before the
peak TKE values as previously shown (in Figures 38 to
41). The 90 CAD spacing of the plots denotes the
approximate spacing between the rising and falling
velocity points in the restrictor over a cycle, or 1/8 of a
cycle. The relative lowest areas of TKE (as shown in
Figures 34 to 37) are around 90 CAD before the
crankangle location for the TKE peaks. During the low
TKE points, the Mach number is rapidly rising.

16
It is interesting to observe that low pressure pulses
arrive at the diffuser throat and generate large shocks,
but at the same time flow separation and TKE decrease
in magnitude. Conversely, when high pressure pulses
reach the throat, throat velocities decrease, but flow
separation and TKE increase. This is somewhat counter-
intuitive as high flow losses are generally associated
with high velocities.
Figure 38. Turbulent Kinetic Energy for the 7 Restrictor
Along Center Plane at 515 Crankangle Degrees
Figure 39. Turbulent Kinetic Energy for the 5.5
Restrictor Along Center Plane at 515 Crankangle
Degrees
The plots for peak and minimum TKE show that in
general the major zone of TKE moves to different areas
of the diffuser over the engine cycle. Looking at the peak
TKE plots (Figures 38 to 41) notice that the highest TKE
values tend to be in the upstream section of the diffuser.
Looking at the low TKE plots (Figures 42 to 45), when
the velocity is highest or rising, the area of the diffuser
showing the highest TKE (at that moment) moves
downstream in the restrictor, but generally has smaller
values.
Also, comparing Figures 38 to 41, the general shape of
the TKE is very similar in the 5.5, 4, and 3, with the 7
being the exception. The cycle low TKE plots in Figures
42 to 45 also exhibit very similar TKE pattern, but again
with the exception of the 7 restrictor.
Figure 40. Turbulent Kinetic Energy for the 4 Restrictor
Along the Center Plane at 540 Crankangle Degrees
Figure 41. Turbulent Kinetic Energy for the 3 Restrictor
Along Center Plane at 615 Crankangle Degrees
Figure 42. Turbulent Kinetic Energy for the 7
Restrictor Along Center Plane at 425 Crankangle
Degrees
17
Figure 43. Turbulent Kinetic Energy for the 5.5
Restrictor Along Center Plane at 425 Crankangle
Degrees
Figure 44. Turbulent Kinetic Energy for the 4
restrictor Along Center Plane at 450 Crankangle
Degrees
Figure 45. Turbulent Kinetic Energy for the 3
Restrictor Along Center Plane at 525 Crankangle
Degrees


TOTAL PRESSURE ALONG THE RESTRICTOR
Pressure recovery through the diffuser helps to give
insight into the efficiency of one design over another.
Researchers often look at static pressure recovery when
determining an intakes flow losses. However the
authors believe total pressure recovery will give a better
idea of restrictor performance and be a better indicator
of losses. Unlike static pressure recovery, total pressure
recovery does not exclude dynamic pressure from the
calculations. However static pressure is also shown for
comparison. A 1.0 bar ambient condition is specified in
the engine model. This results in a 100 kPa total
pressure at the restrictor inlet. Figures 46 to 49 show
pictures of the time averaged total pressure for each
restrictor half-angle at 14,000 RPM.
Again, as witnessed in earlier plots, the 5.5, 4, and 3
half-angle restrictors show a general symmetry about
the main flow axis of the restrictor. However, the 7 half-
angle restrictor shows obvious asymmetry which is due
to a higher a degree of swirling and movement of the
flow throughout the cycle.
Figure 46. Cycle Time Averaged Total Pressure for the
7 Restrictor at 14,000 RPM
Figure 47. Cycle Time Averaged Total Pressure for the
5.5 Restrictor at 14,000 RPM
18
Figure 48. Cycle Time Averaged Total Pressure for the
4 Restrictor at 14,000 RPM

Figure 49. Cycle Time Averaged Total Pressure for the
3 Restrictor at 14,000 RPM
Figure 50 shows the time averaged planar total pressure
along the diffuser for multiple diffuser half-angles. In
Figure 50, the 7 diffuser shows the lowest time
averaged loss of total pressure immediately downstream
of the restrictor throat, which corresponds to an area
ratio that is very near where the flow is shocking down
after the restrictor throat. The 3 diffuser shows the
largest recovery of total pressure along the length of the
restrictor. At a normalized area of 10, the 3 degree
diffuser recovers about 1.9 kPa over both the 7 and
5.5 restrictors. There is very little change in the time
averaged total pressure after an area ratio of 8.
Figure 51 shows the time averaged planar static
pressure along the diffuser for multiple diffuser half-
angles. The 4 and 3 diffusers achieve the lowest value
of static pressure, which is a result of the 4 and 3
diffusers achieving the highest Mach numbers.
Looking again at Figure 50, another interesting item is
the rise in total pressure occurring (on a time averaged
basis) around a normalized area ratio of around 1.5 to
1.6. Typically a rise in total pressure is not considered
possible for a restricted system unless the Reynolds
number is low (< 100 Re) and the flow is under high
normal stresses. [6, 7] The total pressure rise could not
be explained by this phenomenon at this location, but
may occur in other portions of the intake manifold. The
authors believe that the rise in total pressure, after the
restrictor throat maybe due in part, to the pulsation effect
within the intake system. Energy enters into the system
due to pressure pulses which can cause rises in total
pressure. Figure 5, shown previously, helps to visualize
the pressure waves with the anti-node of the wave
typically appearing immediately downstream of the
throat. The restrictor is not choked 100% of the time,
which means that the wave pulsations reaching the
restrictor throat influence the flow field. The other
possibility is that the total pressure rise is partly a
numeric error, as the total pressure is rising across the
shock.
Planar Mean Total Pressure at 14,000 RPM
(Time Averaged)
78
79
80
81
82
83
84
85
86
87
88
89
90
1 2 3 4 5 6 7 8 9 10
Normalized Area
T
o
t
a
l

P
r
e
s
s
u
r
e

[
k
P
a
]
7 5.5 4 3
Figure 50. Cycle Time Averaged Total Pressure for the
7, 5.5, 4, and 3 Restrictors Related to the
Normalized Area Ratio of the Diffuser at 14,000 RPM
Planar Mean Static Pressure at 14,000 RPM
(Time Averaged)
40
45
50
55
60
65
70
75
80
85
1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0
Normalized Area
S
t
a
t
i
c

P
r
e
s
s
u
r
e

[
k
P
a
]
7 5.5 4 3
Figure 51. Cycle Time Averaged Static Pressure for the
7, 5.5, 4, and 3 Restrictor Half-angle Related to the
Normalized Area Ratio of the Diffuser at 14,000 RPM

Figure 52 and 53 show the total pressure over one
engine cycle at various normalized area ratios. In Figure
52, at 455 CAD, a relative minimum of total pressure is
19
achieved at an area ratio of 1.5. Again there are four
total pressure minimums which occur during the cyclic
peak Mach numbers. As was discussed earlier, 455
CAD is a point in the cycle where the Mach number is at
a relative cycle maximum. At 455 CAD, the peak Mach
number occurs at an area ratio of 1.50, and then shocks
down to a subsonic number of about Mach 0.68 at an
area ratio of 1.70.
Figure 52. Total Pressure at 14,000 RPM Over One
Engine Cycle for the Conical-Spline Intake with 7
Restrictor
Figure 53. Total Pressure at 10,000 RPM Over One
Engine Cycle for the Conical-Spline Intake with 7
Restrictor
In Figure 53 at 455 CAD, one would expect a loss in
total pressure from an area ratio of 1.50 to 1.70.
However, the total pressure rises across the area ratio of
1.50 to 1.70. Several possibilities exist for the rise in
total pressure across the shock. The first is numeric
error causing a rise in total pressure. The second is that
3D data has been planar averaged. Some of the 3D
resolved information is lost. There exists a core flow in
the center of the diffuser and this detail is lost in the
planar averaging. As seen in Figures 46 through 49, and
in previous plots, any one variable is not constant across
a cross section. Planar averaging can also introduce
significant weighting averaging errors. [8, 9]
Furthermore, the crank angle resolved plot cannot show
each and every area ratio. The peak Mach number may
be achieved between the area ratio lines shown. Thirdly,
a rise in total pressure may be caused due to pressure
waves caused by the unsteady pulsed nature of the flow
within the intake manifold. As previously mentioned,
Figure 5 showed a standing wave in the intake plenum
at 14,000 RPM.
As noted earlier in the paper, at 10,000 RPM the 7 half-
angle restrictor exhibits subsonic flow through the throat
over the entire engine cycle. In Figure 53, a rise in total
pressure exists during four points in the engine cycle. At
10,000 RPM there are much smaller changes in Mach
number, yet the total pressure still achieves values
greater than 100 kPa four times during the cycle. Figure
54 shows the time averaged total pressure for the 7 and
4 half-angle restrictors at 10,000 RPM. A familiar dip
and rise in total pressure immediately after the throat is
demonstrated in Figure 54, as was also exhibited in
Figure 50. At 10,000 RPM, this dip and rise in total
pressure after the throat is of smaller magnitude and
occurs closer to the throat, in comparison to 14,000
RPM. However, as there are no shocks occurring at
10,000 RPM, the total pressure rise is not a result of any
shock associated numeric errors. The authors believe
the most likely cause of total pressure rise exhibited at
10,000 RPM is due to pressure waves caused by the
unsteady pulsed nature of the flow within the intake
manifold. This effect is also believed to contribute to the
total pressure rise at 14,000 RPM.
Planar Mean Total Pressure at 10,000 RPM
(Time Averaged)
95.4
95.8
96.2
96.6
97.0
97.4
97.8
98.2
98.6
99.0
1 1.25 1.5 1.75 2 2.25 2.5 2.75 3
Normalized Area
T
o
t
a
l

P
r
e
s
s
u
r
e

[
k
P
a
]
7 4
Figure 54. Cycle Time Averaged Total Pressure for the
7 and 4 Restrictor at 10,000 RPM
Figure 55 is the same plot as Figure 54, but extends the
area ratio out to the diffuser exit to show the total
pressure loss over the entire length. At 10,000 RPM the
restrictors continue to lose total pressure over its entire
length which surprisingly contrasts with a leveling off of
total pressure for the same restrictors at 14,000 RPM.
20
Planar Mean Total Pressure at 10,000 RPM
(Time Averaged)
95.4
95.8
96.2
96.6
97.0
97.4
97.8
98.2
98.6
99.0
1 2 3 4 5 6 7 8 9 10 11 12 13
Normalized Area
T
o
t
a
l

P
r
e
s
s
u
r
e

[
k
P
a
]
7 4
Figure 55. Cycle Time Averaged Total Pressure for the
7 and 4 Restrictor at 10,000 RPM
The total pressure graphs showed a higher pressure
recovery along the diffuser for the 3 diffuser and this is
supported by looking at the lower levels of turbulent
kinetic energy along the length of the diffuser. This
detailed level of analysis of determining cause and effect
of changes in diffuser half-angle would not be possible
without the use of coupled simulation methods.
IMPROVEMENT OF RESTRICTOR
PERFORMANCE USING ACOUSTIC DEVICES IN
WAVE
The authors were able to show periods of time that the
pressure waves themselves, disturb the velocity through
the throat. Figures 19 to 25 show the velocity at or near
the throat over one engine cycle, from a coupled 1D/3D
simulation. Table 4 shows that the throat is not choked
100% of the time even at an engine speed of 14,000
RPM. This means that there is time within the engine
cycle that may be possibly exploited by the engine
designers.
ACOUSTIC FILTERING TO REDUCE VELOCITY
FLUCTUATIONS IN THE RESTRICTOR As RPM
increases so does the overall flow. As the overall flow
increases, the restrictor becomes a larger source of
losses. It should also be noted that there are four
periods during the cycle of high velocity, which
correspond to the intake events of the four cylinders
over one cycle. The high velocity portions of the cycle
exist due to a high instantaneous pressure ratio. If the
pressure wave amplitude reaching the throat can be
reduced at a high RPM without negatively effecting the
tuning of the intake, a reduction in peak velocities will
result. Overall a steadier velocity through the throat is
considered desirable to obtain a higher mass flow rate.
There is also a possible side benefit that reducing
pressure and velocity fluctuations might reduce time
averaged turbulent kinetic energy levels in the diffuser.
Acoustic Filtering to Absorb Pressure Pulses To
reduce the pressure fluctuations, and thus the velocity
fluctuations, the authors have investigated the use of
various acoustic filters downstream of the throat to
reduce pressure fluctuations and thus maintain a more
steady velocity. An acoustic filter can absorb some of
the pressure wave energy transmitted into the plenum
from the intake valves. The frequency of each acoustic
filter can be tailored to the RPM range of interest. The
acoustic frequency can be related to the RPM point of
interest by the following formula, Equation 7.
30
*#

=
K
Cylinders of Order Engine RPM
f (7)
Where f is the acoustic frequency and K is a constant,
that equals the number of strokes per engine cycle.
Thus K equals 4 for a four-stroke engine. So for
example, in the engine investigated in this paper, if
attenuation is desired at 13,000 RPM and for the first
engine order, the frequency to be targeted is calculated
as 433.3 Hz.
WAVE was used with meshed manifolds created using
Ricardos WAVEBuild3D. Mesh cell size was set at 15-
20 mm. A 20 mm mesh cell size allows 6 elements per
wavelength up to 12th order at 14,000 RPM. (Note: 6
th

E.O. = 12
th
order for this engine).
Use of a Helmholtz Resonator for Acoustic Filtering A
Helmholtz resonator was chosen to attenuate the
pressure pulses at the target frequency. Other
researchers have found Helmholtz resonators useful for
reducing noise or improving the performance of an
intake manifolds and gas turbines. [10-13] Using a single
orifice allows simpler tuning of the resonator and also
allows more flexibility in placing the orifice. An example
of a Helmholtz resonator is given below in Figure 56.

Figure 56. Example of a Helmholtz Resonator
The equation for the resonance of a one degree of
freedom Helmholtz resonator is given below in Equation
8. [14]
L V
A c
f

=
2
(8)
Where c , is the speed of sound, A is the area of the
resonator throat, L is the length of the throat, and V is
21
the volume of the resonator. An example of the
application of a Helmholtz resonator to the intake
manifold is show in Figure 57. The Helmholtz resonator
was placed under the intake manifold with a single
orifice. The single orifice is located equidistant from all of
the intake runners. The authors believed that locating
any geometric change asymmetric to the runner entry
locations might have a negative impact on cylinder-to-
cylinder volumetric efficiency balance.
The Helmholtz resonator shown in Figure 57 consists of
the following specifications; volume= 0.761 liters, orifice
diameter = 57 mm, and orifice length = 4 mm. At an air
temperature of 300 K, c = 347m/s. This resonator is
referred to as Helmholtz Resonator A. The geometry of
this resonator equates to a tuned frequency of 442Hz or
a tuned RPM speed of 13,260 RPM for the 1
st
Engine
Order. Also, the 7 restrictor was used in this case.
Figure 57. Helmholtz Resonator A Attached to
Underside of Intake
Figure 58 shows the effect of the Helmholtz resonator on
volumetric efficiency in the higher RPM range. Helmholtz
Resonator A shows an increase in VE above the target
RPM for acoustic filtering, in comparison to the baseline
intake. VE has increased in the high rev range of 12,500
to 15,000 RPM. However VE has also increased from
around 10,000 RPM to 12,000 RPM.
Helmholtz Resonator B differs from A in that B has a
slightly larger volume of 0.856 liters. This gives
Helmholtz Resonator B a tuned RPM target of 12,480
RPM for the 1
st
Engine Order. The objective was to shift
the tuned target RPM to a lower engine speed.
However, it can be seen from Figure 58 that the same
VE increase from around 10,000-12,000 RPM occurs,
but the higher RPM increase appears different.
Resonator A shows a bump in VE centered around
13,500 RPM relative to both the baseline intake and the
Helmholtz B intake.
Both Helmholtz intakes show an increase in VE from
10,000 to 15,000 RPM. Both Helmholtz intakes also shift
the peak VE point down in the RPM band. As the use of
a Helmholtz resonator does change the intake volume,
this explains some of the increase in VE in high rpm
band, and the slight downward RPM shift of the VE
peak. These characteristic are similar to trends shown
previously in the paper, due to plenum volume
increases. However, Helmholtz A appears to be
providing an additional benefit in VE around 13,500
RPM even though it does not have the largest intake
volume.
Figure 58. Volumetric Efficiency for Baseline Intake
Along with Two Versions of Helmholtz Resonators
Added to Intake
For Resonator A compared to the baseline, the authors
desired to get confirmation from the analysis that
acoustic filtering was really the means by which the
volumetric efficiency was increasing the most, which
was at 13,750 RPM. Figure 59 shows the reduction in
pressure peaks that occurred by implementing a
Helmholtz resonator as an acoustic filter. An FFT of the
pressure trace showed that only 1
st
Engine Order
content was significantly reduced. However there still
existed pressure fluctuations from other higher order
content.
Figure 60 shows the effect Helmholtz Resonator A had
on Mach number. From the plot, the intake with
Helmholtz Resonator A can be seen to have a smaller
change in Mach number. Referring to Figure 59, it
should be noted that the Mach number peaks occur
when the pressure in Figure 59 is at the lowest. Again,
this demonstrates the velocity in the throat being a
function of instantaneous pressure ratio across the
throat. The popular and discrete view of the flow being
strictly fully choked or not choked is inaccurate. From
Figure 60 it can also be noted that the intake with
Helmholtz Resonator A exhibits a higher percentage of
engine cycle choked, as compared to the baseline
intake.
22
Figure 59. Static Pressure within the Diffuser, 25 mm
Downstream of the Throat, for Helmholtz Resonator A
and Baseline Intake
Figure 60. Mach Number 5 mm Downstream of
Restrictor Throat for Baseline and Helmholtz
Resonator_A Intakes
While the use of acoustic filtering provides gains in a
relatively narrow RPM band, the concept may be more
useful in an engine with a lower cylinder count. As the
cylinder count decreases, flow becomes more unsteady.
The ability to filter out primary firing frequencies with
acoustic filtering may show larger gains in a single or
twin cylinder restricted engine.
Use of an In-line Helmholtz Resonator to Pressurize
Restrictor Inlet A Helmholtz resonator can be used to
attenuate a particular frequency when used as a side
branch as seen in the previous section. However a
Helmholtz resonator inline with the intake system can
provide a frequency-dependent increase in pressure. In
effect the whole intake system can act as a multiple
degree of freedom Helmholtz resonator in itself.
Even if a restrictor is fully 100% choked, mass flow can
be increased by increasing upstream density. [4] The
intent was to add a volume before the restrictor to act as
an inline Helmholtz resonator. As has been discussed
previously, even at high engine speeds (14,000 RPM)
there still exists some wave pressure pulse energy
escaping to the upstream section of the restrictor inlet.
This wave pulsation energy is needed to be able to drive
the Helmholtz resonator. Figure 61 shows an example
(from WAVE) of the application of a volume upstream of
the restrictor inlet.
Two examples of Helmholtz resonators placed upstream
of the throat are examined. Resonator 1 was tuned for
13,400 RPM and Resonator 2 was tuned for 12,550
RPM. The effect of these two different resonators placed
upstream of the restrictor throat is shown in Figure 62.
Figure 61. In-line Helmholtz Resonator Upstream of
Restrictor
Figure 62. Effect of In-line Helmholtz Resonator
Upstream of Restrictor on Volumetric Efficiency
Both cases using a Helmholtz resonator before the
restrictor inlet showed an overall loss in VE when
compared with the baseline run. However, Helmholtz
23
resonator 2 does show a relative bump in VE slightly
above the tuned target of 13,400 RPM. In this case, any
pressure gains at a particular frequency are offset by
losses across the RPM band.
INVESTIGATION OF A FLOW CONTROL DEVICE
USING WAVE-VECTIS
To reduce the onset of separation a flow control device
was placed upstream of the throat in the inlet section.
The device uses four swirl vanes to induce a swirl in the
flow (i.e. induce a tangential velocity component) about
the main restrictor axis. Inducing swirl is intended to
impart radial momentum to the flow downstream of the
throat.
SWIRL GENERATION UPSTREAM OF THE THROAT
Previous work by Gaiser et al [15], showed inducing
swirl into the flow by the use of turning vanes upstream
of a diffuser can reduce separation losses, with gains in
the diffuser outweighing the losses from the vanes.
The Conical-Spline Intake was used with the swirl vane
flow control device. The 7 diffuser was chosen to
investigate the swirl vane flow control device. The
authors chose the 7 diffuser as it had the lowest total
pressure recovery among the various diffuser angles.
The authors wanted to determine if the swirl vane
device could help even the worst performer. Figure 63
shows a sectioned view of the swirl vanes in the area
upstream of the restrictor throat. The swirl vanes are
intended to cause a swirl in the counterclockwise
direction. The swirl vanes start with no slope relative to
the flow, with the slope increasing in the downstream
direction.
Figure 63. Sectioned View of Swirl Vanes Upstream of
Restrictor Throat
A fully coupled WAVE-VECTIS simulation was used to
investigate the swirl vane flow control device. The
baseline intake and the swirl vane intake were both run
at 14,000 RPM. The mesh for the intake with the swirl
vane used a smaller mesh cell (1.75 mm) in the
upstream section of the restrictor than the than the cell
size used in the intake without the swirl vane (3.5 mm). ,
Both intakes used 1.75 mm cells in the diffuser and
throat section as used in previously discussed coupled
runs. The smaller cell in the upstream section was used
to better capture the detail of the swirl vanes. VE
differences between the two runs were negligible. The
intake with the swirl vane device showed 87.8% VE,
while the intake without the swirl vane device showed
87.7% VE. Differences in TKE, total pressure, and Mach
number within the diffuser were also negligible.
The swirl vane did induce swirl with an angular velocity
profile carrying to the throat. However, within 5-10 mm
downstream of the throat, angular velocity was quickly
dissipated. No established pattern of angular velocity
could be distinguished between the baseline intake and
the swirl vane intake.
Using swirl vanes upstream before the throat may
provide some benefit but the angular momentum of the
swirling flow would have to be drastically increased to
determine if there is any real potential benefit. The
authors feel that it maybe worthwhile for future
researchers to investigate more aggressive swirl vanes
to determine if any benefits are possible.

CONCLUSION
The authors research was conducted with Ricardos
Computer Aided Engineering (CAE) software tools,
specifically Ricardo WAVE and VECTIS using 1D and
coupled 1D/3D methods. 1D simulations were carried
out throughout the RPM range whereas the 1D/3D
simulations were only tested at discrete RPM points of
interest. Previous work by the authors found the Conical-
Spline Intake shape to be the best starting point for
achieving high levels of performance.
WAVE results showed that the distance from the
restrictor throat to the intake plenum bottom as caused
by different diffuser half-angles had a large impact on
volumetric efficiency especially from 10,000 to 13,000
RPM. When diffuser lengths were held constant the
differences in volumetric efficiency were smaller. Large
increases in intake plenum volume only resulted in small
increases in volumetric efficiency.
Coupled simulation indicates that 3 half-angle diffuser
provides higher VE to that of the 7, 5.5, and 4 half-
angle diffusers. The 3 half-angle diffuser resulted in a
4.0% improvement in volumetric efficiency over the 7 at
14,000 RPM. The flow structure of the 5.5, 4, and 3
diffuser half-angles appeared to be very similar and
symmetrical along the diffuser. The flow structure of the
7 differed greatly from the other diffuser angles and was
asymmetrical along the diffuser. The authors speculate
that the greater length and packaging difficulties caused
by the use of a 3 half-angle diffuser could probably be
24
overcome by a transition of the diffuser exit to the main
plenum volume at a diffuser area ratio of around 8. The
authors found very little difference in total pressure (and
static pressure) at area ratios greater than 8.
Above 12,000 RPM, WAVE results tended to predict
values of VE greater than the theoretical analysis based
on compressible fluid flow equations and simplifying
assumptions. WAVE-VECTIS prediction gave much
better agreement with theoretical analysis. In some
cases WAVE-VECTIS VE predictions were slightly
greater than theoretical prediction. This is believed to be
caused by the difference in the definition of the VE used
between analysis and simulation. In the analysis,
volumetric efficiency was defined from the reference of
the throat. In the simulation and in practice it is the flow
through the valves in the cylinder that is measured when
calculating the volumetric efficiency.
It has also been shown that while various factors can
determine restrictor performance and how it interacts
with the engine tuning, that the diffuser angle is the
largest factor in effecting volumetric efficiency gains,
even at mid-range RPM.
The restrictor is not choked 100% of the time, even at
high engine speeds. The percentage of the cycle choked
varied with RPM and with the diffuser half-angle. At
10,000 RPM the 7deg restrictor was choked 0% of the
cycle, while the 4deg restrictor was choked 15.3% of the
cycle. At 14,000 RPM the amount of the time the throat
was choked varied from 47.9% to 83.3% of the complete
cycle. Percentage of cycle choked increased as diffuser
half-angle decreased at both 10,000 and 14,000 RPM.
Increased percentage of cycle choked correlated with
increased VE in all instances. Tighter diffuser angles
also had a more uniform Mach number distribution at a
given cross-section than wider angles.
Tighter diffuser half-angles had lower levels of turbulent
kinetic energy and smaller loss of total pressure.
Turbulent kinetic energy peaks shortly after the time at
which velocities in the diffuser peak. Rises in total
pressure are believed to be caused by large pressure
pulses and also due to errors introduced by the use of
area averaging method in VECTIS when applied to non-
uniform flows. Total pressure rise due to numeric error
seems less likely as pressure rise occurred even at
subsonic velocities downstream of the throat.
The use of acoustic filtering methods using Helmholtz
resonators to make flow at the restrictor throat more
uniform and regular appeared promising. A Helmholtz
resonator inside the plenum damped velocity
fluctuations at the throat and offered improvements in
VE at high engine speeds. VE increased at the tuned
rpm point for one intake by approximately 2% over the
expected increase provided by just the additional
plenum volume. A Helmholtz resonator placed inline and
upstream of the restrictor throat showed VE losses at
mid to high RPM. A disadvantage of the resonators is
that their benefit is highly sensitive to differences in
geometry and tuned RPM point.
The use of swirl inducing vanes upstream of the
restrictor throat provided no benefit in terms of
volumetric efficiency improvement. Any swirl generated
by the vanes was quickly dampened out downstream of
the throat. However, more aggressive swirl vanes might
prove more promising and should be considered in
future research.

ACKNOWLEDGMENTS
Special thanks go to our original team advisor Dr.
Patrick Starr for initially sponsoring the research work
and to Dr. David Kittleson and Dr. William Durfee for
continuing their support and sponsorship of our
research. The authors also thank Garrett Stockburger for
his advice and support. The authors would also like to
acknowledge the continual assistance of Dr. Birali
Runesha and the rest of the support staff at the
Minnesota Supercomputing Institute for their help in high
performance computing. We would like to thank Ricardo
Software and in particular Karl John, Patrick Niven, and
Enrico Bradamante for their generous support and
advice. We would also like to thank Kim Lyon of
Daimler-Chrysler for his expertise and advice regarding
engine modeling and calibration. Additional thanks goes
to Dr. Bruce Jones of Mankato State University for
allowing us to use his departments flow bench. We
would like to thank team members John Kess and Aaron
Reinhart for helping us make test fixtures and gather
data for the teams WAVE models.
REFERENCES
1. Society of Automotive Engineers, 2006 Formula
SAE Rules,
http://www.sae.org/students/fsaerules.pdf (31 May
2006).
2. Claywell, M. R., Horkheimer, D. P., and Stockburger,
G. R., Investigation of Intake Concepts for Formula
SAE Four-Cylinder Engine Using 1D/3D (Ricardo
WAVE-VECTIS) Coupled Modeling Techniques,
SAE 2006 Motorsports Conference, 2006-01-3652.
3. Guizzetti, M. and Colombo, T. Combined WAVE-
VECTIS Simulation of an Intake Manifold of V6 PFI
Gasoline Engine,
http://www.ricardo.com/download/pdf/VECTIS_comb
ined_wave_VECTIS.pdf (31 May 2006).
4. Saad, M. A., Compressible Fluid Flow, 2
nd
Edition,
Prentice Hall, 1993, ISBN 0-13-161373-1.
5. Bressler, H., Rammoser, D., Neumaier, H., Terres,
F., Experimental and Predictive Investigation of a
Close-Coupled Catalytic Converter with Pulsating
Flow, SAE 960564.
25
6. Issa, R. I., Rise of Total Pressure in Frictional
Flow, AIAA Journal 1995, Vol. 33, No. 4, pp. 772-
774.
7. Homann, F., The Effect of High Viscosity on the
Flow Around a Cylinder and Around A Sphere,
NACA-TM-1334.
8. Wyatt, D. D., Analysis of Errors Introduced By
Several Methods of Weighting Non-Uniform Duct
Flow, NACA-TN-3400, 1955.
9. Pianko, M. and Wazelt, F., Suitable Averaging
Techniques in Non-Uniform Internal Flow,
Propulsion and Energetics Panel Working Group 14,
AGARD-AR-182, 1983.
10. Dupre, I. D. J., and Dowling, A. P., The Use of
Helmholtz Resonators in a Practical Combustor,
Journal of Engineering for Gas Turbines and Power,
April 2005, Vol. 127, pp. 268-275.
11. Onorati, A., Montenegro, G., and Errico, G. D.,
Prediction of the Attenuation Characteristics of I.C.
Engine Silencers by 1-D and Multi-D Simulation
Models, SAE 2006-01-1541.
12. Wan, D., and Soedel, D. T., Two Degree of
Freedom Helmholtz Resonator Analysis, SAE 2004-
01-0387.
13. Selamet, A., Dickey, N. S., Radavich, P. M., and
Novak, J. M., Theoretical, Computational and
Experimental Investigation of Helmholtz Resonators:
One-Dimensional Versus Multi-Dimensional
Approach, SAE 940612.
14. Winterborne, D., and Pearson, R., Design
Technologies for Engine Manifolds Wave Action
Methods for IC Engines, 1999.
15. Gaiser, G., Oesterle, J., Braun, J., and Zacke, P.,
The Progressive Spin Inlet Homogenous Flow
Distributions Under Stringent Conditions, SAE
2003-01-0840.


CONTACT
Mark R. Claywell has a Bachelor in Mechanical
Engineering from the University of Minnesota.
clay0052@umn.edu
Donald P. Horkheimer has a Bachelor in Aerospace
Engineering and Mechanics from the University of
Minnesota.hork0004@umn.edu
http://www.me.umn.edu/~hork0004

DEFINITIONS, ACRONYMS, ABBREVIATIONS
CAD: Computer Aided Design or Crankangle Degrees
CAE: Computer Aided Engineering
CFD: Computational Fluid Dynamics
FFT: Fast Fourier Transform
RPM: Revolutions per Minute
TKE: Turbulent Kinetic Energy
VE: Volumetric Efficiency

NOMENCLATURE
General
A Area [m
2
]
x
A Area at Distance from Throat [m
2
]

A Area at Restrictor Throat [m


2
]
c Speed of Sound [m/s]
f Frequency [Hz]
Ratio of Specific Heats
(

v
p
C
C

K Engine Stroke Constant [4]
L Length [m]
Arbitrary Scalar -
Flow Uniformity Index -
P Pressure [Pa]
Density [kg/m
3
]
Q
&
Volumetric Flow Rate [m
3
/s]
V Volume [m
3
]
Subscripts
0
Stagnation Conditions
i
Scalar at a Mesh Cell in a Plane
mean
Average Scalar Value across a Plane

Vous aimerez peut-être aussi