You are on page 1of 12

1

2006-01-0044
Electromagnetic Fully Flexible Valve Actuator
David Cope and Andrew Wright
Engineering Matters, Inc.
Copyright 2006 SAE International
ABSTRACT
An electromagnetic fully flexible valve actuator
(FFVA) for internal combustion engines is described
which offers the potential for significant improvements in
fuel economy, emissions, and performance, especially at
low end torque, in internal combustion engines. The
FFVA offers variable lift and timing combined with
controllable seating velocity. It operates on a design
principle distinct from existing actuators: the
electromagnetic actuator exerts appreciable bi-
directional force throughout the device stroke mitigating
the need for mechanical spring-derived resonance. The
FFVA is a direct drive device with a unique magnetic
structure that combines high bandwidth and strong
forces to meet the engine performance requirements.
This paper presents the innovative electromagnetic
design, simulation, and bench testing of the actuator on
a single cylinder engine.

INTRODUCTION
Variable valve actuator (VVA) strategies have been
proposed for decades as a means for improving
performance and efficiency while controlling emissions
[1,2]. The benefits to engine performance of various
valve actuation strategies, such as Early Intake Valve
Closing (EIVC), Late Intake Valve Closing (LIVC), Late
Intake Valve Opening (LIVO), and Variable Max Valve
Lift (VMVL), have been thoroughly investigated and
experimentally verified. These strategies range from
cylinder deactivation to discrete step and continuously
variable cam profiling, and, ultimately, to camless
technologies such as electrohydraulic and
electromagnetic actuation [3,4,5,6]. However, despite
the performance benefits most current internal
combustion engines do not take advantage of VVA.
Those that do are extensions of standard cam
technologies [1,3,7,8].
Camless technologies have the capability to fulfill
the promise of FFVA. Specifically, the ability to fully
define the motion of the engine valves combined with
intelligent control enables the adoption of any valve
actuation strategy achieving the above mentioned
benefits. Electrohydraulic valve actuation has a long
history as a research tool for quickly varying cam profiles
to study valve lift vs. timing and is technically capable of
achieving all the requirements of FFVA. Recent efforts
to apply electrohydraulic valve actuation to production
engines have focused on reducing power consumption
as well as redesigning prohibitively expensive
components such as the high pressure pump [6].
Electromagnetic valve actuation potentially achieves the
requirements of FFVA while avoiding the complexity of
an additional hydraulic system. The potential barriers to
the FFVA adoption are increased electrical power
consumption, too great a valve seating velocity,
unacceptable actuator failure modes, cost, and actuator
packaging difficulties. Additionally, current cam valve
train technology has evolved to an extremely well-
developed and thoroughly tested system setting a high
standard for replacement technologies [3].
The objective of the current research is to develop
an actuator capable of satisfying the FFVA concept. A
new, innovative, patented, and patent pending
electromagnetic valve actuator for internal combustion
engines is discussed. Fully flexible valve actuation is
achieved through concurrent design of electromagnetic,
electrical, mechanical, and thermal aspects. This
actuator achieves fully flexible valve actuation through
variable valve timing as well as variable lift. Among the
numerous advantages of FFVA are increased engine
efficiency over the engine speed and load range, and
the elimination of the cam and throttle subsystems.
Demanding power, force, speed, and control
requirements have prevented standard actuators from
fulfilling FFVA principles, prompting the development of
highly complex mechanical and electromechanical
systems. The innovative actuator discussed herein
consists of stationary permanent magnets, a stationary
coil, and moving iron stem that transmits bi-directional
forces to the valve. Under contract to the National
Science Foundation, an experimental proof-of-principle
actuator was developed and mounted on a single
cylinder engine. Experimental data confirm basic
operational capabilities, with both variable timing and lift.
The projected system power requirements for a 16 valve
system are low enough to be operable from a standard
12V automotive electrical system with alternator
augmentation [9,10,11].
Magnetic finite element analysis and simulation of
the actuator demonstrated the achievement of the
desired performance objectives (power, force, and
control) for a FFVA.
2
A traditional cam drive train, shown in Figure 1, acts
on the valve stems to open and close the valves. As the
crankshaft drives the camshaft through gears or a belt,
the timing of the opening and closing of a valve is
controlled by the cam design and is fixed relative to the
piston position. This means that the engine
performance is optimized only over a narrow range of
engine speed and load. Most existing electromagnetic
valve actuators focus on variable timing (the ability to
open and close the valves at will) however, they do not
allow for variable lift. The FFVA approach, shown in
Figure 2, allows for fully flexible valve control (i.e. both
variable timing and variable lift, low valve seating
velocity, fast transition times (up to 6000 rpm), and full
stroke force authority).
Electromagnetically-controlled valves can operate
optimally at all engine speeds, torque levels, and
temperatures thereby greatly improving the engine
performance, including emissions. For example,
improvement in fuel economy in excess of 15% is
expected for FFVAs [12, 13].

Figure 1. Traditional Mechanical Cam Drive Train

Figure 2. First Generation Fully Flexible Valve Actuator


VALVE ACTUATOR STRATEGIES
With the additional degrees of freedom offered by
FFVA (lift, timing, and seating velocity) several
strategies for manipulating the intake flow exist. The
most attractive strategies are those which eliminate the
need for intake throttling and in the process, reduce
pumping losses. Specifically, EIVC, LIVO, LIVC, and
VMVL are the general intake valve actuator strategies
which achieve this [3]. A qualitative comparison by
Sellnau of the performance of these four strategies
reveals EIVC to be the most favorable strategy for
reducing pumping loss and LIVO to provide the best
mixture motion at ignition. The other two strategies,
LIVC and VMVL, employ only one of the allowed
degrees of freedom and are shown to be generally less
effective by comparison. LIVC changes only the
duration of the intake opening process, while VMVL
varies only the lift. The other two strategies, EIVC and
LIVO may be accomplished with variable timing alone,
but are preferentially combined with lower lift.
Two complementary methods for electromagnetic
valve actuation exist: commanded holding and
commanded acceleration. The holding style actuator
relies on stored mechanical potential energy to be
released into kinetic energy for transition, then storing
once again the kinetic energy into mechanical potential
energy to be held until the next desired transition. The
acceleration style actuator electrically supplies the
kinetic energy, and then reclaims the kinetic energy
through regeneration. In theory, both methods require
only the initial energy input for operation. In detail,
however, the power requirements and achievable
dynamics of each method differ based upon the specific
actuator properties such as force per current over
stroke.
In a practical embodiment, most actuators are a
hybrid of these two methods. An example of the holding
style actuator, the Pischinger design, uses solenoids to
hold the valve either in the fully open or fully closed
position while mechanical springs are in full compression
[14]. When the solenoid releases the armature, the valve
travels through the spring equilibrium point then
decelerates to the end of the travel at which time
another solenoid holds the armature in place. In this
way, electromagnetic force is used only to hold the
solenoid and spring force is used to accelerate/
decelerate the valve. In order to achieve the fast
transition times required at high RPM, the spring
constant is tuned to a resonance of the combined mass
of the armature and valve. The selection of solenoid
actuation is well suited to this method because of the
one-way actuation (holding) and the limited stroke range
of appreciable force. The disadvantage of this method is
the dependence on resonant transition allows only
variations in the timing, but not the lift or speed of
transition. While holding force alone is adequate for
actuation, the inability to push the armature requires
complex control for soft valve seating [15].
In contrast, the commanded acceleration method,
with force authority over the entire stroke, is capable of
3
controlling the valve lift directly. A bidirectional force
approach is also better suited to controllable valve
seating. Potential difficulties with this method are
obtaining enough force for high speed valve acceleration
and maintaining reasonable power requirements. The
benefits of the commanded acceleration method include
the capability of achieving the valve actuation strategies
mentioned above utilizing variable valve lift, variable
timing, and seating velocity control.

FFVA DESIGN
The primary thrust of the actuator design is to use
the best combination of methods discussed above,
holding and accelerating, which will achieve the optimal
valve strategies with reasonable power requirements
and robust construction. The continuing development for
fully flexible valve actuation is an iterative design
process with the following stages: magnetic finite
element analysis combined with parameter optimization
routines, power minimization through a tailored current
profile, thermal analysis, dynamic simulation, and failure
mode effects and analysis.

ACTUATOR CONFIGURATIONS
A common issue for actuator design is which
components move relative to an external body and
which other components are stationary. Typically there
are two possibilities: (1) moving magnets (MM) and
stationary coil, or (2) moving coils (MC) and stationary
magnets. As discussed in the following sections, we
have developed another option, (3) moving plunger (MP)
with stationary coil and magnets.
Typical operation of an engine might average 2000-
3000 rpm for 15,000 miles/year, which equates to
approximately 30-45 million actuation cycles/year. A MC
configuration has potential issues with flexing of
electrical leads, and a MM configuration has potential
issues with magnet mechanical damage (fatigue or
cracking) due to the constantly reversing acceleration
profile. Therefore, a configuration without either of these
components moving is indeed attractive. A MP
configuration can be made extremely rugged and able to
withstand the high acceleration cycles. Furthermore, the
stationary magnets and stationary coil can be
electrically, thermally and mechanically buffered to some
extent within their environment. As will be seen, this is
the approach used for the FFVA.

MAGNETIC FINITE ELEMENT ANALYSES (MFEA)
As mentioned above, the central concept of the
actuator permits several quite different configurations
[20,21]. Commercial magnetic finite element analysis
software (Maxwell

3-D from Ansoft Corp. [16]) was


utilized to input the various configurations and analyze
the performance of each configuration. In addition,
using the Optimetrics component, we performed
optimization of each configuration. The central idea of
Optimetrics is an automated way to numerically
determine the sensitivity of a design configuration to a
defined goodness parameter and then to continue to
refine the design in the direction of increasing
goodness. When the design space is large (many
independent parameters), there are simply too many
possible permutations of design variables to compute
each one individually. For example, there are more than
fifteen individual dimensional characteristics which
collectively, with materials choices and boundary
conditions, define a single design. If 6 values of each
dimension were to be analyzed, over 11 million design
combinations would result for each major configuration.
This is too many to seriously evaluate and would
represent analyst overload. Instead, Optimetrics
essentially calculates the greatest slope toward
increasing goodness and marches along that path.
Localized extremes were encountered and dealt with by
utilizing multiple starting designs to verify convergence
to an optimized specific design combination.
Nine design configurations were examined. The
eight unselected designs are shown in Figure 4 through
Figure 11 in Table 1 and the selected design is shown in
Figure 12. (Figure 3 provides a key for interpretation of
the symbols in Table 1.) The configurations are
compared based upon the metric of the ratio of valve
acceleration and square root of dissipated input
power, P A/ . The best performing configuration,
Figure 12, has the greatest value of this metric. The
valve acceleration is computed by computing the
electromagnetic force and dividing it by the sum of
moving masses of the valve, the valve stem, and
connecting hardware. [Note: other bases for
comparison could, of course, be used. In that case, the
optimized actuator design details would differ.] The
MFEA simulation takes current density in the coil as an
input and calculates the coil resistance and electrical
power consumed. When configurations can be
employed as either MM or MC, the reported figure is for
the underlined configuration. There are, of course,
many parameters upon which the value of P A/
depends, and the utility of the actuator is not described
solely by this metric. Therefore, although we considered
heavily the value of P A/ in determining which
configuration to build for the feasibility demonstration,
we also considered aspects such as expected actuator
reliability and longevity. Based upon this analysis, the
final FFVA design is markedly superior to the other
designs. Fundamentally this is because the moving
portion is simply a steel plunger of relatively small radius
and relatively high magnetic saturation value. This
design, once properly engineered for the environment, is
expected to have an excellent reliability record because
of its simple, robust nature. In addition, by use of the
reluctance forces, the actuator can be designed to have
the valve closed during power off. This will greatly
reduce potential valve-piston interference events.
4


Figure 3. Key for Table 1.

Table 1. FFVA Configurations Analyzed and Optimized with Magnetic Finite Element Software; Axis of Symmetry is shown
dot-dashed on the left of each figure.



Coil
Coil
C
L

Figure 4. Either MM
or MC configuration.
P A =39.1
sec
1
kg




Figure 5. Preferred MC,
but could be MM
(magnets-plus-iron
move); the iron mass is a
significant penalty.
P A =45.6
sec
1
kg







Figure 6. Either
MM or MC
configuration.
P A =28.5
sec
1
kg
Coil
Coil
Coil

Figure 7. Either MM or MC
configuration. Three coils
consume significant electrical
power but add substantial force
capability.
P A =29.1
sec
1
kg


Figure 8. MC
configuration: iron
moves with the coil
and focuses the field.
P A =50.7
sec
1
kg


Figure 9. MC
configuration: iron
moves with the coil and
focuses the field.
P A =59.4
sec
1
kg


Figure 10. MC
configuration.
Connecting to valve
would be a
challenge.
P A =40.8
sec
1
kg

Figure 11. MC configuration:
iron moves with the coil.
P A =60.0
sec
1
kg

5

Figure 12. Selected FFVA Design for the
demonstration engine. = P A/ 75
sec
1
kg

0 20 40 60 80
Figure 4
Figure 5
Figure 6
Figure 7
Figure 8
Figure 9
Figure 10
Figure 11
Figure 12
Acceleration/Sqrt(Power) [1/sqrt(kg*sec)]

Figure 13. Comparison of FFVA Configurations:
Valve acceleration per square root of dissipated
power.

The preferred FFVA configuration is shown in Figure
12. Figure 13 shows a chart comparing the predicted
values of P A . The selected design has a
substantially greater figure of merit than the other
designs. Figure 14 shows the FFVA selected design in
RZ symmetry. Figure 15 shows the magnetic field
vectors of the design to accelerate the valve in a
downward direction. Figure 16 shows the magnetic flux
lines for this scenario.
Figure 17 shows the electromagnetic force on the
plunger (connected to the valve via the stem). It is seen
that there is a force present even when there is no
current (middle curve). This is due to the magnetic
reluctance force of the ferromagnetic plunger in the
permanent magnet structure. Essentially, the plunger
tends to move upward if it is above the centerline or
tends to move downward if it is below the centerline; it
would remain in either of the extreme upper or lower
positions. Since this curve of force over stroke with no
current is approximately linear with distance, it can be
nearly cancelled by appropriate choice of mechanical
spring. The other two curves in the figure represent the
electromagnetic forces associated with applying positive
or negative current, respectively, to the actuator. Figure
18 shows the results of using a linear mechanical spring
to cancel the reluctance force. Note the force is not
quite constant over stroke and there is a slightly greater
force available at the beginning and end of a valve
transition to accelerate and decelerate the valve,
respectively.
Fe Fe Coil
Axis of
Symmetry

Figure 14. Electromagnetic FFVA Geometry in R-Z
Symmetry

6

Figure 15. FFVA Magnetic Field Vectors in R-Z
Symmetry


Figure 16. FFVA Magnetic Flux lines in R-Z Symmetry


-200
-100
0
100
200
-4 -2 0 2 4
Positive Current
Zero Current
Negative Current
Stroke, mm
F
o
r
c
e
,

N

Figure 17. Hysteretic FFVA Forces vs. Stroke and
Current, uncompensated (raw) electromagnetic forces.
-200
-100
0
100
200
-4 -2 0 2 4
Positive Current
Zero Current, reluctance force
Negative Current
Decreasing Stroke
Increasing Stroke
Stroke, mm
F
o
r
c
e
,

N

Figure 18. Hysteretic FFVA Forces vs. Stroke and
Current, compensated by a linear spring.

OPTIMUM DYNAMIC ACCELERATION PROFILE
In order to accomplish the rapid valve transitions
necessary for high speed valve actuation (closed to
open, and open to close), various valve acceleration
profiles were investigated. Increasing the applied
current provides a greater acceleration force, but also
increases the dissipated electrical power. It was desired
to discover the acceleration profile which achieved the
specified dynamic performance (essentially moving a
fixed distance, d, in a time, t) at minimum electrical
dissipated energy. High speed transition dynamics
require valve motion of 8 mm in 3.3 ms. A symmetric
acceleration/ deceleration profile requires a motion of
d
0
=4 mm in t
0
=1.65 ms. An actuator force linearly
proportional to the current was assumed. Figure 18
shows the force is not constant over the stroke but is
greatest when accelerating the valve; for present
purpose we assume a constant
I
F
k = (force per
current) is approximately achieved by a reluctance-
compensated FFVA actuator. This assumption yields
acceleration,
m
t I k
x a
) (
= = & & , so


=
0
0
0
) (
t
dt
m
t I k
dt d .
The dissipated energy to be minimized is

0
0
2
) (
t
dt R t I
and we desire to find the waveform ). (t I This
minimization problem is subject to the constraint that the
valve moves the specified distance in the specified time.
It can be solved by calculus of variations and Lagrangian
multipliers or by trial and error. In either event, it can be
readily verified that a current waveform of

=
0
1
1 ) (
t
t
I t I with an acceleration profile of

=
0
1
1 ) (
t
t
a t a , where
7

= = =

=
2
0
0
0 2
0
0 1
1
2
2
3
2
3 3
t
d
a
t
d
m
I k
a , meets the
constraints while minimizing the energy.
0
a is the
constant value of acceleration required to move a
distance d
0
in time t
0
. In fact, ) (t a provides an energy
savings of 25% compared to a constant acceleration
profile (for a constant ratio of force and current).
Therefore, this is the desired acceleration waveform for
the FFVA. Note, however, that this optimized waveform
requires an increase in instantaneous force of 50%
compared to a constant acceleration profile.

0
150
300
450
0 0.5 1.0 1.5 2.0 1.65
Optimized Linear Profile
Constant Acceleration Profile
Time, ms
A
c
c
e
l
e
r
a
t
i
o
n
,

g
e
e

Figure 19. Ramped and constant acceleration profiles for
the first half of the opening transition.

0
1
2
3
4
0 0.5 1.0 1.5 2.0 1.65
Optimized Linear Acceleration
Constant Acceleration Profile
Time, ms
V
a
l
v
e

D
i
s
p
l
a
c
e
m
e
n
t
,

m
m

Figure 20. Valve displacement vs. time for the two
acceleration profiles for the first half of the opening
transition.

0
0.1
0.2
0.3
0 0.5 1.0 1.5 2.0 1.65
Difference in Valve Displacement
Time, ms
D
i
s
p
l
a
c
e
m
e
n
t
,

m
m

Figure 21. Difference in valve displacement for the two
profiles. The ramped acceleration profile opens the valve
more quickly than the constant acceleration profile.

THERMAL ANALYSIS
Most high performance electrical machines are
ultimately thermally limited. This is because electrical
machines, especially permanent magnet machines,
improve in performance with increases in excitation
current. Therefore, the common practice is to increase
the current until either portions of the design
magnetically saturate, or the local temperature increase
required to transport the electrically dissipated power is
unacceptably high. Since the neodymium-iron-boron
magnets themselves have a maximum operating
temperature of 150C, they frequently provide the lowest
upper limit to allowable energy dissipation. Figure 22
shows the geometry analyzed. A steady state
temperature plot is shown in Figure 23 for an excitation
of 80W continuous and boundary conditions of forced
convection to 100C.
Note the annular heat pipes between magnets
appear to increase the magnet temperature locally, but
they transport heat from the coil and ultimately reduce
the magnet temperature by 5-10C. Heat flux vectors
are shown in Figure 24. As indicated in the figure,
roughly one-quarter of the heat leaves the actuator
through the top, one-quarter through the bottom, and
one-half of the heat exits through the outer diameter of
the device.
Fe Fe Coil
Annular
heat pipes
Annular
heat pipes

Figure 22. Electromagnetic FFVA Geometry in R-Z
Symmetry. The 4 annular heat pipes help cool the coil.
8


Figure 23. FFVA Steady State Temperature Plot showing
the magnets do not exceed 135C.


Figure 24. FFVA Heat Flux Q. The annular heat pipes
between magnets transport significant thermal power.

DYNAMIC SIMULATION
The block diagram for the FFVA simulation is shown
in Figure 25. Briefly it can be described as follows: Initial
valve position is determined and based upon an engine
map and a current command is issued to move the valve
to a position. The coil current is converted to force by
the number of coil turns and look-up tables created from
MFEA output, which account for hysteresis of valve
position (i.e., reluctance forces) and desired direction of
motion. The dynamic equations (F=ma, F = sum of
external spring forces, etc.) are then computed to obtain
valve acceleration, velocity and position. Force-per-
current data derived from Figure 18 is compiled into a
look-up table to provide accurate force and power
characteristics over the full stroke. The simulation has
been utilized to simulate different acceleration profiles
and failure modes and effects. Friction work of 100 mJ
per transition was taken into account.


Figure 25. Simplified FFVA Simulation Block Diagram

The most stressing dynamic scenario for the
actuator is for high speed valve transitions. Therefore,
simulations for 6000 rpm are provided below.
Figure 26 shows the forces required for three
simulated acceleration profiles: Constant acceleration;
Ramped acceleration; and Ramped acceleration with
spring-back. Constant acceleration is the simplest case
and allows for the minimum force to provide the required
dynamics. Ramped acceleration is the linear decrease
in acceleration with time for minimum dissipated power.
Ramped acceleration with spring-back is the linear
decrease in acceleration for the first half of the opening
transition, followed by valve free-flight for the second
half of the transition, followed by a stiff spring encounter
which compresses and expands to reverse the valve
velocity, finally, the valve is gradually decelerated to a
low velocity seating. The rebound spring significantly
reduces dissipated power for high speed operation and
allows low lift operation at lower speeds since it does not
contact the valve below approximately 8 mm stroke.
Note that the constant acceleration profile swings
between equal and opposite acceleration values.
Ramped acceleration is the linear decrease in
acceleration with time, followed by a linear increase in
acceleration with time. Ramped acceleration with
spring-back is the linear decrease in acceleration in
combination with a stiff spring at the extent of the valve
opening. This utilizes a slightly-reduced peak force.
Computed valve motion resulting from these three
acceleration profiles is shown in Figure 27. It is seen
that the constant acceleration profile provides the
slowest opening characteristic while the Ramped
acceleration with spring-back may involve valve over-
9
travel beyond the nominal 8 mm stroke. Alternatively,
the spring-back can be designed to occur at 7.5mm-
8mm and so not increase total valve displacement. This
would actually reduce the dynamic requirements on the
actuator (reduced force, current, and power).
Based upon the simulation, the total dissipated
electrical and mechanical energy can be calculated for
the different acceleration profiles. The results are shown
in Table 2. In addition to the electrical energy
dissipation, 0.200 J work per cycle (100 mJ per
transition) has been included to account for friction and
other energy loss mechanisms. Note that the Ramped
acceleration profile reduces power consumption by 30%
compared to the Constant acceleration profile, which is
greater than the 25% predicted earlier for a constant
value of k=F/I. This is because, as shown in Figure 18,
the force (and hence force per unit current) is greater for
valve acceleration and hence even more energy is
saved. Basically, the actuator is more efficient at initial
acceleration than at late acceleration. The Ramped
acceleration with spring-back reduces the average
power approximately 50% more since it essentially
eliminates the energy for valve motion reversal.

-300
-200
-100
0
100
200
300
0 1 2 3 4 5 6 7 8 9 10 1.65
Constant acceleration
Ramped acceleration
Ramped acceleration with spring-back
Closing Opening
Time, ms
F
o
r
c
e
,

N

Figure 26. Simulated Required Forces vs. Time for three
possible acceleration profiles (6000 rpm, 3.3 ms
transition).

-1
0
1
2
3
4
5
6
7
8
9
0 1 2 3 4 5 6 7 8 9 10 1.65
Constant acceleration
Ramped acceleration
Ramped acceleration with spring-back
Closing Opening
Time, ms
S
t
r
o
k
e
,

m
m

Figure 27. Simulated Valve Stroke vs Time resulting from
the three acceleration profiles.

Table 2. Cycle energetics (6000 rpm, 3.3 ms transition
time, 20 ms period)
Profile Cycle
Energy
Average
Power
(1 valve)
Average
Power*
(16 valves)
Constant acceleration 3.75 J 188 W 3008 W
Ramped acceleration 2.65 J 133 W 2128 W
Ramped acceleration
with spring-back
1.36 J 68 W 1088 W
*Intake valves only.

FAILURE MODES
A common complaint against camless valve
actuation technologies is the position of the valve after a
failure in the valve actuation system. If the failure is
sudden, then mechanical inertia will continue to drive the
pistons through their trajectory making valve-piston
interference a possibility. In many systems the position
of the valve is indeterminate; while in others the valve
reverts to a position midway between fully open and fully
closed. In either event valve-piston collision is highly
likely, especially in todays high-compression-ratio
engines.
Two features of the FFVA design presented here
dictate the position of the valve in the event of a failure.
As shown in Figure 18, the actuator will preferentially
reside in the closed position due to magnetic reluctance.
Within limits, the reluctance force can be controlled by
initial design, recognizing that there is a trade-off with
acceleration performance. A second feature of the
actuator results from its full stroke force authority. The
coil can be wound two-in-hand meaning that
electrically isolated coils excite the actuator. Hence,
even if an electrical short or open occurs in one of the
subcoils, then the other subcoil still has substantial force
over the valve. Shown in Figure 28 and Figure 29, are
two simulations of sub coil failures at different times.
They show that upon failure of one subcoil, the
remaining working subcoil can be excited to recall the
valve home (closed position) within a cycle before the
piston rises to top dead center (TDC), thereby averting
an interference condition. The working subcoil can hold
the valve in place, aided by the reluctance force. In
Figure 28, the failure occurs at 1.0 ms and the valve
opens 6 mm. In Figure 29, the failure occurs at 1.6 ms
and the valve opens the full 8 mm. A key feature of the
Ramped acceleration (and so ramped force) profile is
evident here: due to the decreasing force profile, most of
the energy has already been imparted to the valve for
the transition by the time the failure has occurred.
Other failure modes are certainly possible and these
are presently being investigated.

10
-50
0
50
100
150
200
250
0 2 4 6 8 10 12 14 16 18 20
-2
0
2
4
6
8
10
Force
Valve Displacement
Piston arrives at TDC
Valve parked home and held Closed
Valve attracted home
with Subcoil #2
Usual
Closed
time
Failure occurs
in Subcoil #1
Time, ms
F
o
r
c
e
,

N
S
t
r
o
k
e
,

m
m

Figure 28. Simulated failure of Subcoil #1 at t=1.0 ms.
Subcoil #2 has sufficient attractive force to retrieve and
park the valve Closed, thereby eliminating valve-piston
interference.

-50
0
50
100
150
200
250
0 2 4 6 8 10 12 14 16 18 20
-2
-1
0
1
2
3
4
5
6
7
8
9
10
Force
Valve Displacement
Piston arrives at TDC
Failure occurs
in Subcoil #1
Usual
Closed
time
Valve attracted home
with Subcoil #2
Valve parked home and held closed
Time, ms
F
o
r
c
e
,

N
S
t
r
o
k
e
,

m
m

Figure 29. Simulated failure of Subcoil #1 at t=1.6 ms.
Similar to the above figure, but the failure occurs very late
in the transition. Subcoil #2 has sufficient attractive force
to retrieve and park the valve Closed, thereby eliminating
valve-piston interference.


SINGLE CYLINDER ENGINE DEMONSTRATION
ACTUATOR CONSTRUCTION AND ENGINE
PREPARATION
After optimizing the configuration and geometry of
the actuator and running dynamic simulations, an intake
valve actuator was fabricated and installed on a single-
cylinder engine. (The exhaust valve cam was retained.)
A drawing of the single-cylinder engine selected for this
demonstration project is shown in Figure 30. Selection
criteria for the engine were: relatively low power (4-hp),
light weight, and ease of access to the valves. This
engine was particularly easy to modify the valve since it
has both overhead cams and overhead valves.

EXPERIMENTAL RESULTS
Feasibility of operating an internal combustion
engine based upon the designed electromagnetic valve
actuator was demonstrated. An increasingly difficult
series of actuator tests consisted of operating the
actuator under the following conditions: on a bench top,
in the unassembled cylinder head, statically on the
engine (assembled), slow motion on the engine (hand
crank), fast motion on the engine (pull-cord operated),
and during combustion.

SINGLE CYLINDER ENGINE OPERATION
Operation of an engine with the actuator controlling
the intake valve was the major goal of this feasibility
demonstration project. Figure 31 shows the valve
displacement during the cycle as a function of crank
angle from a throttled no-load run at 1500 rpm. Twin
desirable attributes of high opening speed (~890
mm/sec) and low landing speed (~30 mm/sec) are
evident in the figure. It should be remarked that the
actuator operated the first time the engine was started
and the engine runs reliably, although changing engine
speeds requires a manual adjustment in valve
commands. Figure 32 and
Figure 33 show the actuator mounted on the engine.
Note details of the setup and diagnostics are evident in
the figures.


Figure 30. Honda GC135QHA Engine [22]

0
1
2
3
4
0 90 180 270 360 450 540 630 720
Test #4
Opening speed
~890 mm/sec
Landing speed
~30 mm/sec
Crank Angle, deg
L
i
f
t
,

m
m

Figure 31. Experimental Valve Displacement vs Crank
Shaft Angle
11

Figure 32. Phantom Views of Cam Magnet Path

Figure 33. FFVA mounted on Engine


FUTURE WORK
The FFVA development effort is on-going under a
National Science Foundation grant. Continuing
improvements in magnetic configuration will influence
the other aspects of the design cycle: thermal, electrical,
and mechanical. In addition to refining the concurrent
design parameters, an in depth study of valve control will
be pursued. The current state of knowledge in variable
valve control strategies is well developed and largely
applicable to the developing actuator [17, 18, 19].
Further development in control is vital to realize the full
capabilities of the actuator system. A position estimator
of some kind is certainly necessary. We presently use a
position sensor, but since the device is fundamentally a
permanent magnet machine, sensorless control is
entirely possible, eliminating a component of cost and
failure modes. Detailed simulations will be performed to
investigate integrated valve actuation strategies and
control, as well as testing potential failure modes and
effects. The NSF grant will culminate with fabrication
and installation of the FFVA system on an automotive
engine. The valve and engine performance will be tested
with dynamometer and gas analyzer to measure the
FFVA-equipped engine performance and emissions.
Commercial feasibility of the actuator has not been
demonstrated. With the decreasing cost of permanent
magnets and the low cost of the remaining iron and
copper components, the high-volume cost estimate of
the actuator may approach the value of the eliminated
and removed mechanical components.

SUMMARY & CONCLUSIONS
The major objectives of this research were to
design, build, and test an electromagnetic fully flexible
valve actuator. These objectives were achieved. The
design process involved many configurations with
moving magnets, moving coils, and/or a moving plunger.
The selected design performed significantly better than
other designs based on dynamic performance
(acceleration per square root of dissipated power) and
had desirable characteristics such as stationary magnets
and stationary coil.
A dynamic simulation was created to predict
performance of the valve under a variety of conditions,
especially various acceleration profiles. Valve-piston
interference is avoidable during failure modes.
Detailed manufacturing drawings were made and an
actuator was fabricated and assembled. A single-
cylinder engine was chosen and modifications were
made to the cylinder head to mount the actuator and
various sensors. A series of run-up tests was performed
culminating in the feasibility demonstration of engine
operation under electromagnetic intake valve actuator
control. In conclusion, the results show the actuator
exhibits the inherent advantages of the fully flexible
valve actuator such as variable timing, variable lift, and
low valve-landing speed.

ACKNOWLEDGMENTS
The authors would like to thank Dr. J im Cowart of
MIT and the U.S. Naval Academy, Dr. Christopher
Corcoran of Corcoran Engineering, and Mr. David
Fischer of DMF Associates for their input, support, and
professional advice. The authors would also like to
thank MIT graduate student Mr. Bernard Yen for his help
with the experimental work. The authors acknowledge
the support of the National Science Foundation through
its SBIR program.

12
REFERENCES
1. Stone, R., and Kwan, E., Variable Valve Actuation
Mechanisms and the Potential for their Application,
SAE Paper 890673, 1989.
2. Klein, F. et al, The Influence of the Valve Stroke
Design in Variable Valve Timing Systems on Load
Cycle, Mixture Formation and the Combustion
Process in Conjunction with Throttle-free Load
Governing, SAE Paper 981030, 1998.
3. Sellnau, M., and Rask, E., Two-Step Variable Valve
Actuation for Fuel Economy, Emissions, and
Performance, SAE Paper 2003-01-0029, 2003.
4. Ahmad, T., Theobald, M., A Survey of Variable-
Valve-Actuation Technology, SAE Paper 891674,
1989.
5. Dresner, T., Barken, P., A Review and
Classification of Variable Valve Timing
Mechanisms, SAE Paper 890674, 1989.
6. Schechter, M., and Levin, M., Camless Engine,
SAE Paper 960581, 1996.
7. Pierik, R., Burkhard, J ., Design and Development of
a Mechanical Variable Valve Actuation System,
SAE Paper 2000-01-3307, 2000.
8. Kreuter, P. et al, The Meta VVH System A
Continuously Variable Valve Timing System, SAE
Paper 980765, 1998.
9. Rivas, J ., Perreault, D., and Keim T., Performance
Improvement of Alternators With Switched-Mode
Rectifiers, IEEE Trans. Energy Conv., Vol. 19, No.
3, Sept. 2004, p. 561
10. Tang, S.C., Keim, T., and Perreault, D.J ., Thermal
Modeling of Lundell Alternators, IEEE Trans.
Energy Conv., Vol. 20, No. 1, March 2005, p. 25.
11. Liang, F., Miller, J . M., and Xu, X., A Vehicle
Electric Power Generation System with Improved
Output Power and Efficiency, IEEE Trans Ind.
Appl., Vol. 35, No. 6, Nov./Dec. 1999, p. 1341.
12. Carney, D. "Internal Combustion Engineering,"
Automotive Engineering International, May 2004, p.
42;
13. "European Centers of Power," Automotive
Engineering International, J une 2004, p. 67.
14. Pischinger F., Kreuter, P., "Arrangement for
Electromagnetically Operated Actuators", U.S.
Patent #4,515,343, May 7, 1985, www.uspto.gov..
15. Wang, Y., Peterson, K., et al, Modeling and Control
of Electromechanical Valve Actuator, SAE Paper
2002-01-1106, 2002.
16. Ansoft Corp., Pittsburgh, PA 412-261-3200,
www.ansoft.com.
17. Ashhab, M., Stefanopoulou, A., Camless Engine
Control for a Robust Unthrottled Operation, SAE
Paper 981031, 1998.
18. Ashhab, M-S; and Stefanopoulou, A., Control-
Oriented Model for Camless Intake Process Part
I, Transactions of the ASME Vol 122, March 2000.
19. Ashhab, M-S; and Stefanopoulou, A., Control of a
Camless Intake Process Part II, ASME J ournal of
Dynamic Systems, Measurement, and Control
March 2000.
20. Wright, A., Cope, D., High Intensity Radial Field
Magnetic Array and Actuator, US Patent
#6,876,284, April 5, 2005, www.uspto.gov.
21. Wright, A., Cope, D., High Intensity Radial Field
Magnetic Array and Actuator, US Patent #6,828,890,
December 7, 2004, www.uspto.gov.
22. Honda Engines Owner's Manual GC135/GC160,
Honda Motor Co., Ltd, 2001.

CONTACT
The authors may be contacted at:

Engineering Matters, Inc.
375 Elliot St., Suite 130K
Newton, MA 02464
617-965-8974

Dr. David Cope: dcope@engineeringmatters.com
Andrew Wright: awright@engineeeringmatters.com.

NOMENCLATURE
P A/
Ratio of Acceleration to the square root of
Power; used as a metric of valve
performance
EIVC Early Intake Valve Closing (valve actuation
strategy)
EVA Electromagnetic Valve Actuator
FFVA Fully Flexible Valve Actuator
LIVC Late Intake Valve Closing (valve actuation
strategy)
LIVO Late Intake Valve Opening (valve actuation
strategy)
MC Moving Coil
MFEA Magnetic Finite Element Analysis
MM Moving Magnet
MP Moving Plunger
R-Z
Symmetry
Axisymmetric cross-section
VMVL Variable Max Valve Lift (valve actuation
strategy)
VVA Variable Valve Actuator

You might also like