Sei sulla pagina 1di 11

A High Performance, Continuously Variable Engine

Intake Manifold

2011-01-0420
Published
04/12/2011

Adam Vaughan and George J. Delagrammatikas


Cooper Union
Copyright 2011 SAE International
doi:10.4271/2011-01-0420

ABSTRACT
Manifold tuning has long been considered a critical facet of
engine design and performance optimization. This paper
details the design, analysis and preliminary testing of a
continuously variable, carbon fiber intake manifold for a
restricted 2003 Suzuki GSXR-600 engine. The device
achieves a large dynamic runner length range of 216-325 mm
through the use of a half-tube, sliding shell design that differs
substantially from traditional variable intake approaches. A
combination of Ricardo WAVE and 2D/3D Ansys Fluent
simulations were used to aid in the design of the intake along
with a custom software routine to optimize restrictor
geometry through fully automated CFD simulations. The
sliding mechanism was actuated via a cable linkage system
and powered by a small servo motor. This motor was
controlled by a Microchip dsPIC microcontroller that was
embedded in a custom power distribution PCB for the 2009
Cooper Union Formula SAE entry. The controller
communicates with the engine's MicroSquirt ECU over
CAN to read instantaneous engine speed and commands the
servo based on an empirically tuned look-up table. Initial
testing of the intake showed the expected torque and power
variation, maintaining over 95% of the peak engine torque for
an additional 60% of the usable engine speed range in
addition to a peak power improvement of 5% relative to a
baseline static intake configuration. A peak power
improvement of over 22% was also achieved relative to the
2008 FSAE intake. The variable intake system adds less
than 1% to overall powertrain weight and is able to actuate
the full dynamic range in less than 1.0 s. Additional gains are
expected through optimized cam timing coupled with
refinements to the initial engine calibration.

INTRODUCTION
As is characteristic in any rotary power-plant (electric,
internal combustion, etc.), there exist various operating points
where torque output reaches a local or global maximum [1].
Because torque affects the acceleratory performance of a
vehicle, optimizing it is a critical factor for any vehicle
designer. For engines, torque is strongly affected by
Volumetric Efficiency (VE), which is a measure of the actual
air inducted versus the swept volume of the piston [2]. VE, in
turn, is closely coupled to the resonance conditions that
develop in the engine's manifolds, the timing of the induction
and exhaust processes, fluid flow losses and the average
speed of fluid flow [1].
Many empirical and computational studies have shown a
clear relationship between the geometry of manifold ducts
and volumes to the resonance peaks favorable to high VE
performance (see Figure 1) [1]. As a result, an engine
manifold designer is able to tune an engine to achieve
higher VE at certain engine speeds by carefully selecting
appropriate geometry. However, drivability also remains a
concern, and because highly tuned engines typically exhibit
narrow operating bands of peak performance, it is the goal of
this paper to explore dynamically varying manifold geometry
in an effort to produce an engine that is tuned over a greater
range of engine speeds. By taking this approach, a
performance engine can be made more drivable (through a
more constant, or flatter torque curve) without a
corresponding increase in engine displacement to compensate
for low torque regimes.

Engine displacement must not exceed 610 cc.


Any component capable of throttling the engine must be
upstream of the flow restriction.
All intake components must be contained within an
envelope defined by the main roll hoop and rear tires.

Figure 1. Generalized intake geometry for a restricted


FSAE engine
A flatter torque curve is especially important in the
displacement limited Formula SAE competition, which is
the target application for the design to be presented. It is
worth mentioning that modern Variable Valve Timing (VVT)
systems can also be utilized to produce a more drivable
vehicle, with greater torque gains as well as benefits in
reduced exhaust emissions and improved fuel economy [3]
[4]. However, a variable intake is more easily adapted to the
small, fixed cam timing motorcycle engines commonly used
in FSAE. It is also simpler in overall construction.
Numerous models have been developed to help predict VE
curves as a function of runner length and other manifold
geometry [2]. These range from a basic mass-spring model
(Helmholtz) [2], to 1D pressure wave simulations (such as
Ricardo WAVE) [5], to even transient 3D CFD codes that
strive to model the complete fluid state (such as Ansys
Fluent and Ricardo VECTIS) [6] [7] [8]. Each scheme
has its own advantages and disadvantages. The Helmholtz
model, for instance, is a simple equation that roughly predicts
the VE peak but does not yield information about other
operating points [2]. While the Helmholtz model is
convenient for calculations, it can show a sizable error at high
speed engines [9].
Ricardo WAVE discretizes duct lengths and plenums as 1D
flow elements to in order to rapidly compute a fairly accurate
VE curve, but it struggles to capture effects produced by
turbulence (such as from the restrictor [5] [7] [8]). Ansys
Fluent and Ricardo VECTIS are each capable of
providing a more complete model compared to Ricardo
WAVE (as they include empirically verified turbulence
models), but they do so at great computational expense [7]
[8].
As mentioned earlier, this paper will only focus on the
FSAE class of vehicles and will, therefore, be compliant
with the rules that govern the 2010 competition. The key
requirements for the competition are:
There must be a 20 mm diameter flow restriction on all air
entering the engine.

It is worth briefly commenting on the difficulties that arise


from the FSAE constraints before delving into greater detail
in the subsequent sections. The 20 mm flow restriction
presents what is probably the greatest modeling challenge.
Not only does it cause the engine to always operate at part
load conditions, but it also necessitates a modeling approach
that can handle compressible, transient flow as well as
turbulence. Thus, it is important to incorporate elements of
3D simulations to effectively model this element. The
FSAE displacement limitation constrains off the shelf
engine choices to those of the small engine motorcycle/ATV
market. Finally, the envelope limitation presents a handful of
system packaging issues, particularly with the long restrictor.

DESIGN
The general design methodology taken for this study begins
first with restrictor optimization, followed by a discussion of
plenum sizing and finally the selection of variable runner
length geometry through 1D and 3D CFD codes and
packaging limitations. Although not presented, an embedded
intake servo control system was developed for this project
using a Microchip dsPIC30F4011/30I microcontroller. This
system utilized a Controller Area Network (CAN) interface to
query the MicroSquirt Engine Control Unit (ECU) for
instantaneous engine speed and to subsequently command an
optimal runner length based off an empirically determined
look-up table.

ENGINE SELECTION
Previous Cooper Union FSAE entries have used the 2003
Suzuki GSXR-600 (specifications given in Table 1) engine
primarily because its stock configuration provides more lowend torque compared to other engines within its displacement
class. Of course, adding a restrictor and packaging an
alternate set of manifolds will alter this characteristic [1]. The
engine has proven itself a reliable workhorse, and given the
team's prior experience with this engine as well as the
availability of spare parts within the lab, it was decided to
continue using this power plant. The engine is in stock
configuration with the exception of the ECU, intake manifold
and an increased compression ratio provided by a Wiesco
piston upgrade.

Table 1. A list of general specifications for the 2003


Suzuki GSXR-600 engine

shift the VE resonance peak lower. Directly conflicting these


results, Hamilton et al. empirically found that a large 6 L
plenum increased the peak torque output by 17% [12].
However, it is worth noting that the runner length used for
their study was 150 mm, which is significantly shorter than
the runner lengths detailed in this and other studies. Because
the ultimate goal of this project was to study the effects of a
variable runner length intake and given the highly non-linear
results found by previous studies with large plenums, the
final plenum volume was sized primarily off the space
required for runner actuation (3.1 L). It is left to future
computational and empirical studies to further optimize this
variable.

RESTRICTOR

1D CFD OPTIMIZATION

With one of the key handicaps of FSAE requirements being


the choked flow induced by the 20 mm diameter restriction,
extensive studies were performed to optimize the restrictor
element for maximum mass flow. A custom software routine
was developed in C# to perform an exhaustive Design of
Experiments (DoE) search quantifying the relative
contributions of various geometric variables. The details of
this study are described in [10], however the final selected
geometry is presented in Table 2.

As runner length is a variable that can produce sizable VE


effects [1], a method of varying length was incorporated into
the intake design. To initially explore the feasibility of such a
setup, preliminary calculations were first performed using a
simple Helmholtz model [2]. The model results showed that
in order to cover a substantial range of VE peaks
(8000-12000 rpm), the runner length must more than double.

Table 2. Ranges of variables used for DoE restrictor


study, with final selected values

PLENUM SIZING
Plenum volume is known to affect engine VE [1]. In
unrestricted engine simulations, Winterbone and Pearson
found that small volumes were favorable at low engine
speeds and that plenum effects were marginal at high speeds.
They went on to state that plenum sizing is mostly a concern
of idle speed control. For restricted engines, Blair has
recommended that plenum volume be as large as possible
[11], and up to 6 L has been successfully used by other
FSAE teams without throttle response issues [12].
McKee et al. stated that 3.5 L was the optimal volume for a
600 cc engine from their simulations [11]. Through
simulations of conical spline FSAE intakes, Claywell et al.
found that large plenum volumes (>3.5 L) beneficially
affected VE, but the results were highly dependent on the
way plenum geometry was adjusted to vary the volume [7]
[8]. In general, they found that the gains from increased
plenum volume showed diminishing returns (<1% of VE)
after increasing beyond 5 L. Additionally, they were able to
show that adjusting the shape of a large plenum volume could

In order to further quantify the expected gains of varying


various runner lengths, a Ricardo WAVE model was
developed for the 2003 Suzuki GSXR-600 engine. A rubber
mold of the engine's intake port was made, digitized and
incorporated into the engine model to improve its accuracy.
The rubber mold geometry was also used to aid in injector
placement later on in the design.
A contour plot of the resulting torque at Wide Open Throttle
(WOT) for various runner lengths across different engine
speeds is presented in Figure 2. This chart clearly shows how
varying runner length can shift the engine's torque peak,
allowing the engine to produce more torque over a greater
range of engine speeds. To more clearly show the simulated
gains, a plot of the percent difference relative to the torque
output at a runner length of 320 mm is shown in Figure 3.
While the Ricardo WAVE model requires further
calibration using dynamometer results (data which prior to
this project were unavailable for the restricted Suzuki
GSXR-600), its general trends were found to agree with
those of the Helmholtz model. Note that the simulated torque
gains of 5% shown here were significantly less than the
10-15% found in the McKee et al. empirical study with
similar runner length variation but different geometry than
used here [11]. From the results of this Ricardo WAVE
study and those of [7], [8] and [11], a desired length variation
of 205-330 mm was selected. This maintained high torque
output for a substantial portion of the torque band
(7500-10,000 rpm) and did not unnecessarily increase runner
length to be optimal for engine speeds that did not produce
adequate power. The difficulties of packaging this length

variation will be discussed later, but the shortest packageable


runner (without modifying the engine head) is roughly
indicated in Figure 2.

Dynamic range is further reduced because the housing tube


must devote space for a fuel injector, a bend for restrictor
packaging, and a mounting flange.
While not simulated, the relatively harsh bend (necessary to
increase dynamic range) would most likely induce
recirculation within the head port.

Figure 2. Ricardo WAVE simulated contours of torque


(Nm) for various runner lengths at WOT
Figure 4. Section view of telescoping runner design
After iteratively optimizing the design around these
considerations, the final geometry could vary (at most) from
240-300mm, which significantly fell short of the desired
range of 205-330 mm. Nevertheless, this geometry was
rushed to fabrication in order to have a design ready for the
2009 FSAE competition.

Figure 3. Ricardo WAVE simulated % difference in


torque at WOT from 320mm runner length baseline

METHODS OF RUNNER VARIATION


In order to package the desired runner length variation of
205-320 mm, two approaches were pursued. The first was a
telescoping design (Figure 4), wherein a movable tube (red)
would slide into and out of larger fixed tube (light grey)
much like a trombone (see [10]). While relatively simple, this
design proved less than ideal because of the following
considerations:
The dynamic range is geometrically limited, as the sliding
tube cannot actuate to a length longer than the tube within
which it is housed.

For the 2010 FSAE season, the partially complete


telescoping design was abandoned to pursue a new design
that showed the promise of remedying the dynamic range
issues while building on the fabrication experience developed
during the 2009 FSAE season. This new design (Figure 5)
operated under the principle that only the large area change at
the end of the tube into the plenum served as a wave
reflecting resonance device, and that the geometry after the
end of the tube did not contribute appreciably to the
resonance characteristics. To accomplish length variation,
only half the tube (red) was actuated to increase or decrease
the runner's centerline length, while the other half of the tube
remained stationary and built into the plenum wall.
By taking this approach:
The full dynamic range could be realized.
Harsh bends within the intake could be avoided.
The linkage system could provided by a simple cable
mechanism.

STEADY-STATE 3D CFD
OPTIMIZATION

Figure 5. Overview of half-tube concept


Despite these advantages, the design time for the half-tube
approach was greater due to packaging issues with the long
restrictor and the long (220 mm) sliding surface along which
the half-tubes actuate. These two features together were
difficult to fit around the engine block, the frame and the
FSAE surface envelope constraint while trying to minimize
the number of bends and number of fabricated parts.
Figure 6 shows a detailed section view of the new design
concept. The cross-sectional area of the fixed runner was
carefully designed to have at most 1.5% variation relative to
the intake port area as it transitioned from the engine's oval
geometry to a circular shape. Blair had found that such a
transition was beneficial to VE [6]. Additionally, the variable
runner geometry includes a 25% reduction in inlet crosssectional area for the fully extended runner position which
was sized using unpublished, empirical cross-sectional area
studies done by Purdue FSAE. Reducing runner diameter is
discussed extensively in the literature [1] to improve low
engine speed VE, although the particulars of the approach
described here are unique in that only the end condition area
is reduced. The reduction in area was primarily done to
improve the flow characteristics of the half-tube design, by
providing a more gradual transition within the packaging
constraints detailed above. That said, future studies should
further optimize this transition.

As an exhaustive, transient simulation of manifold dynamics


was not possible within the time constraints of this project, a
steady-state solution was used to help identify areas for
improvement. This was done in an effort to minimize minor
flow losses which have a much more profound effect on VE
at high engine speeds [1]. It is understood that steady-state
solutions do not fully capture the complex manifold dynamics
that develop within the manifold (see [7] and [8] for a more
exhaustive study).
The key areas of interest were:
The flow field leading into the half-tube bellmouth.
How the flow distribution varied as the runners were
actuated between the various runner positions.
If recirculation developed in the diffuser region after the
restrictor terminated at the top of the plenum.
Whether there were any localized regions of Mach >0.5
outside the restrictor.
The following discussion will display contour plots along
various planar slices through the combined fluid volume of
the plenum and runners. These slices are necessary to
visualize a 3D flow field on a 2D piece of paper. Figure 7
should be used as a reference as to which cross-sectional slice
is being discussed.

Figure 7. Contours of Mach number for various


rendering planes under 6900 Pa pressure drop (regular
ellipse geometry)

Figure 6. Detailed section view of half-tube concept

The pressure-based 3D solver was used with the realizable turbulence model for these studies [7]. The imposed
pressure drop was 6900 Pa, which was chosen such that the
flow was partially choked. Claywell et al. had determined
that the restrictor is never fully choked, even at an engine
speed of 14,000 rpm, and thus a partially choked condition

was felt to be more reflective of the typical operating regime


in which the intake would be operated [8]. Again, a full
transient simulation is necessary to truly capture the manifold
dynamics.
With regard to Mach number, Winterbone and Pearson
discussed empirical data that intake components should be
sized such that the Mach number does not exceed 0.5,
otherwise a precipitous drop of in VE would ensue [1]. By
this criterion, the FSAE mandated restrictor is clearly the
biggest shortcoming of the designed intake (Figure 7).
The first problem region identified through steady-state 3D
CFD simulation was an area of recirculation within the
original runner design (Figure 8). This issue was not easily
remedied, as the original runner geometry had been designed
to package the intake within the FSAE surface envelope
and other components of the 2009 vehicle. Nevertheless, the
mass flow gains (>8%) from removing this region of
recirculation were enough to justify a review of the original
packaging. After extensive iterative optimization of the
original centerline curve, an arrangement was found that
packaged the intake well around the previously discussed
criteria, with only a marginal (<10 mm) increase in fixed
runner length.

Figure 9. Top view of velocity contours along midplenum plane


The simple radius proved to be a localized point of turbulence
generation, and is included in Figure 9 as a baseline for
comparison for later intake geometries. The reader should
note that the maximum mass flow is appreciably lower in the
simple radius design primarily due to recirculation in the
runners (see Figure 8) which was remedied in the later
designs.
In order to explore the mass flow variation through the entire
dynamic range of runner lengths, an additional steady-state
study was performed to identify fluid flow losses as the
intake was actuated from its shortest to longest position. The
results of the study are presented in Table 3. These
simulations were done using the regular ellipse geometry
(except where elongated ellipse geometry is noted).
Table 3. Steady-state mass flow variation for various
runner positions

Figure 8. Velocity vectors along the mid-runner plane


for original and final runner geometry
Blair identified the optimal bellmouth geometry to be a short,
wide ellipse (the parameters of which are best described
within [6]). He came to this conclusion using transient, 3D
simulations of simple radii, airfoils and ellipse geometries
with differing lengths and widths. In the interest of
comparing the relative contribution of the bellmouth
geometry on peak mass flow for the designed intake, a series
of studies were performed using the largest ellipse geometry
that could package well (Figure 9).

From these data, it is clear that the outer two cylinders do not
receive as much flow at long runner lengths, as they are offset
from the centerline of the restrictor. This may have been
remedied through a longer diffuser section after the restrictor,
but packaging and fabrication constraints did not afford
enough space to accomplish this. Also shown is that the
elongated ellipse geometry reduced the cylinder imbalance
from 2.8% to 2.3% for the 90 mm runner length.

In addition to mass flow studies, the inlet to the runner for the
final geometry was carefully inspected using a 3D vector
view of the flow field. It was found to have a smooth
transition into the runner, with no significant points of
recirculation or other anomalies. This suggests that the
method of runner length actuation itself does not present
appreciable flow losses while still allowing the full dynamic
range of runner lengths.

MECHANICAL DESIGN
Hobby servo motors, linear actuators and solenoids were
screened for their potential use in actuating the variable
runners. It was found that solenoids and linear actuators were
unnecessarily heavy for this application and required a
linkage system to cover the full dynamic range. High-end
hobby servos did not present these issues, and thus an
actuator scheme was implemented around a high-torque
Futaba S5801 sail winch. This servo provided 6 full
rotations (which represents significantly more than the typical
servo rotation of 180 degrees) and it came stock with a winch
drum for winding up cables. This drum proved to be of an
appropriate size for actuating the full dynamic range of the
variable runners. To efficiently package the system, a mount
was incorporated into the fuel rail for the servo and cable.
One of advantages to the half-tube intake concept was that a
simple, low-tolerance cable linkage setup could slide the
runners along the bottom plenum surface. Additionally, by
using a cable linkage, the servo could be placed external to
the intake without a complicated sealing method. The reasons
this was advantageous were two-fold: For one, the servo was
perceived to be the most frequently serviced component and
easy access to it was desirable. The second reason was that it
minimized the quantity of small parts residing within the
plenum. In the event of an intake failure, such small
components could easily make their way into the cylinder and
cause engine damage. The basic actuation mechanism is
shown in Figure 10. It consists of a hardened stainless steel
shaft and two high-misalignment linear bearings that guide
the variable runners.
To seal the variable runners, o-ring stock was installed in
channels along the bottom of the variable runners (see Figure
6). With the application of lubricant, having the o-ring stock
ride along the surface of the bottom plenum offered a simple
solution to sealing the runners as they actuated. Additionally,
this approach is able to take advantage of the negative
dynamic pressure developed within the runner from the mean
flow into the cylinders. To seal the rear portion of the
variable runner, flaps of rubber were incorporated and glued
onto the back side of the fixed runner (see Figure 6). The
final selected geometry is shown in Figure 11. Note that only
a single fixed runner is displayed to simplify the diagram.

Figure 10. The sliding mechanism

Figure 11. Final design overview


The material of choice for this low production run, weight
sensitive application, with thin walls and complex curves is
Carbon Fiber Reinforced Polymer (CFRP). While the
particulars of CFRP design and fabrication are outside the
scope of this paper, a quasi-isotropic layup schedule was used
to aid in FEA simulation and because of the expected uniform
pressure loading. It was necessary to use a coring material
(Divinycell H100 Semi-rigid PVC foam) to enhance the
strength of the predominantly flat plenum surfaces. A
geometrically stronger curved plenum design where the
variable runners rotated was considered, but the fabrication of
the design was deemed too complex. A detail view of the
fabricated intake actuation system is shown in Figure 12.
After assembly, timed trials of the actuation system showed
that the intake was able to transition its full dynamic range
(shortest to longest runner length) in under 1.0 s. The static
friction resistance to sliding with lubrication was measured to
be 8 N. Figure 13 shows the fabricated intake manifold on the
engine prior to the installation of the top plenum half.

PRELIMINARY RESULTS
The fabricated intake was installed on a recently rebuilt 2003
Suzuki GSXR-600 (specifications given in Table 1). The
engine had completed a break-in and calibration procedure
using the 2008 intake. For the new intake, the engine was
recalibrated for maximum torque at WOT with the runner
length actuated to the shortest runner position.

Figure 12. Detail view of fabricated half-tube concept

After calibration, the engine was allowed to slowly sweep


through a speed range of 7000 to 12000 rpm (at WOT) while
torque was logged. This was repeated twice for 10 different
runner lengths. The two runs at each runner length were
averaged and all of the results are plotted in Figure 14. Note
that the torque data for each sweep were collected after the
engine's internal gearbox using a Mustang MDT-70 eddy
current dynamometer and corrected using SAE J1349
compensation. Pump gasoline (93 AKI) was used for all runs
and an Innovate Motorsports LC-1 wideband Air-Fuel
Ratio (AFR) sensor was used for closed-loop fueling control.

Figure 14. Contours of measured torque (Nm) at WOT


Although a direct measurement of VE would have more
accurately separated intake performance from combustion
and friction considerations, it was not available for these
tests. For these preliminary trials this was deemed acceptable
provided fueling was adjusted under closed-loop AFR
control. Because engine torque is proportional to VE when all
other variables are constant [2], these data do provide a rough
indication of the VE variation. Figure 15 presents the percent
difference in torque from a 325 mm runner length baseline
(similar to Figure 3).
Figure 13. Fabricated intake on the engine prior to
installation of the top plenum half

through a continuously variable system. In addition to being


variable, these results also demonstrate a peak power output
improvement of greater than 22% versus the 2008 intake
design.

Figure 15. % difference in measured torque at WOT


from 325 mm runner length baseline
The results show the expected shift in the resonance peak of
the engine. Compared to the simulations presented earlier, the
minimum torque is significantly higher in value while the
peak torque follows a more complex distribution. Figure 16
shows an envelope of all the torque values presented in
Figure 14. The torque curve when the runner length is equal
to 289 mm is also shown as a baseline for comparison
between fixed and variable operation. This baseline runner
length was selected on the basis that it has a relatively long,
flat torque curve that would be desirable for the driver on a
road course if the intake was operated in a fixed runner length
configuration.

Figure 17. Power curves for selected runner lengths at


WOT
Transient operation of the variable intake under servo control
is presented in Figure 18 for a constant engine speed of 9500
rpm. Since the system is able to respond to a less than 1.0 s
step input, these torque data support the assumption that
closed-loop fuel adjustment is able to compensate for VE
variation. The reader should note that the torque data shown
here are higher in value than those presented earlier. This
discrepancy is due to improvements that were made to runner
sealing after the initial dataset was acquired (Figure 14). The
gradual decay in torque output is due to inadequate cooling
within the test cell. Heat was also identified as a source of a
servo failure, and it will necessitate either relocating the servo
away from the engine or designing a new servo system with
greater temperature tolerance.

Figure 16. Envelope of all torque values at WOT


compared to 289 mm performance
By shifting the runner position, the intake is able to improve
the torque output of the engine by 2-5% relative to the
289mm baseline. Figure 17 shows how the power peak
transitions for selected runner lengths. This plot shows that
the typical variable intake approach of toggling between two
runner lengths misses a sizable portion of the gains to be had

Figure 18. Transient runner length transitions at


constant engine speed (9500 rpm) and WOT

CONCLUSIONS &
RECOMMENDATIONS
The design, analysis and successful testing of a continuously
variable, carbon fiber engine intake manifold has been
described in this work. The intake can vary its runner length
from 216 to 325 mm. Preliminary testing of the intake
showed the expected torque and power variation, with a peak
power improvement of over 22% relative to the 2008 design
and a 5% torque output gain compared to a 289 mm runner
length baseline. This variable intake system increases the
powertrain weight by less than 1% and can transition through
its full dynamic range in less than 1.0 s.
The intake was designed through published data, a Ricardo
WAVE model and steady-state 2D/3D Ansys Fluent CFD
code. Optimization of restrictor geometry was accomplished
through a custom software routine that was developed to
automatically mesh and execute cases in Ansys Fluent [10].
Final intake geometry was determined through a series of
packaging and fabrication compromises detailed in this work.
The variable runner mechanism was actuated through a servo
motor and a cable linkage which was controlled by a
Microchip dsPIC microcontroller.
In the future, restrictor DoE software should be expanded into
a more general automated design optimization package to aid
in more than just restrictor design. An extension has already
been coded to interface to MATLAB for such routines. It is
believed that by limiting trials to good performers from the
current steady-state results and systematically limiting the
selection process with a gradient-based optimization scheme,
the simulations can be accelerated enough to allow a full
transient design optimization. Additionally, the preliminary
performance data should be incorporated into the Ricardo
WAVE model to better calibrate it and allow for better
computer aided engine optimization.
It is hoped that future studies can build on the hardware and
software developed for this project to further improve and
characterize the half-tube variable intake concept. The current
engine calibration should be refined (with optimized cam
timing) and coupled with the empirically determined runner
length versus engine speed look-up table code embodied in
the Microchip dsPIC microcontroller. Controller code
should be expanded to incorporate traction control and
driving style considerations. Finally, a direct measurement of
VE should be used for future studies.

REFERENCES
1. Winterbone, D. E., and Pearson, R. J., Design Techniques
for Engine Manifolds, Professional Engineering Publishing,
London, ISBN 186058179X, 1999.

2. Heywood, J. B., Internal Combustion Engine


Fundamentals, McGraw-Hill, New York, ISBN 0-07-028637X, 1988.
3. Brstle, C. and Schwarzenthal, D., VarioCam Plus - A
Highlight of the Porsche 911 Turbo Engine, SAE Technical
Paper 2001-01-0245, 2001, doi:10.4271/2001-01-0245.
4. Kreuter, P., Heuser, P., Reinicke-Murmann, J., Erz, R. et
al., Variable Valve Actuation - Switchable and Continuously
Variable Valve Lifts, SAE Technical Paper 2003-01-0026,
2003, doi:10.4271/2003-01-0026.
5. Cauchi, J., Farrugia, M., and Balzan, N., Engine
Simulation of a Restricted FSAE Engine, Focusing on
Restrictor Modelling, SAE Technical Paper 2006-01-3651,
2006, doi:10.4271/2006-01-3651.
6. Blair, G. P., Cahoon, W., Special Investigation: Design
of an Intake Bellmouth, Race Engine Technology: 34-41,
September 2006.
7. Claywell, M., Horkheimer, D., and Stockburger, G.,
Investigation of Intake Concepts for a Formula SAE FourCylinder Engine Using 1D/3D (Ricardo WAVE-VECTIS)
Coupled Modeling Techniques, SAE Technical Paper
2006-01-3652, 2006, doi:10.4271/2006-01-3652.
8. Claywell, M. and Horkheimer, D., Improvement of
Intake Restrictor Performance for a Formula SAE Race Car
through 1D and Coupled 1D/3D Analysis Methods, SAE
Technical Paper 2006-01-3654, 2006, doi:
10.4271/2006-01-3654.
9. Zimmerman, S., Cordon, D., Anderson, M., and Beyerlein,
S., Development and Validation of an Impedance Transform
Model for High Speed Engines, SAE Technical Paper
2005-01-3803, 2005, doi:10.4271/2005-01-3803.
10. Vaughan, A. and Delagrammatikas, G.J., Variable
Runner Length Intake Manifold Design: An Interim Progress
Report, SAE Technical Paper 2010-01-1112, 2010, doi:
10.4271/2010-01-1112.
11. McKee, R.H., McCullough, G., Cunningham, G., Taylor,
J.O. et al., Experimental Optimisation of Manifold and
Camshaft Geometries for a Restricted 600cc Four-Cylinder
Four-Stroke Engine, SAE Technical Paper 2006-32-0070,
2006, doi:10.4271/2006-32-0070.
12. Hamilton, L. J., Cowart, J. S., Lee, J. E., and Amorosso,
R. E., The Effects of Intake Plenum Volume on the
Performance of a Small Naturally Aspirated Restricted
Engine, ASME Paper ICEF2009-14036, 2009.

ACKNOWLEDGMENTS
Fabrication of the parts described in this work would not have
been possible without the CNC expertise of Sinisa Janjusevic.
James Abbott's guidance was instrumental with the various
acoustical aspects inherent to this project. Numerous Formula

SAE team members have also assisted with this project:


Muneeb Hai, Greg Shikhman, Rob Smith, Kim Meehan and
Tim Fedullo.

The Engineering Meetings Board has approved this paper for publication. It has
successfully completed SAE's peer review process under the supervision of the session
organizer. This process requires a minimum of three (3) reviews by industry experts.
All rights reserved. No part of this publication may be reproduced, stored in a
retrieval system, or transmitted, in any form or by any means, electronic, mechanical,
photocopying, recording, or otherwise, without the prior written permission of SAE.
ISSN 0148-7191

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.
SAE Customer Service:
Tel: 877-606-7323 (inside USA and Canada)
Tel: 724-776-4970 (outside USA)
Fax: 724-776-0790
Email: CustomerService@sae.org
SAE Web Address: http://www.sae.org
Printed in USA

Potrebbero piacerti anche