You are on page 1of 9

Downloaded from SAE International by Univ of California, Friday, July 27, 2018

2014-01-2222
Published 09/16/2014
Copyright © 2014 SAE International
doi:10.4271/2014-01-2222
saeaero.saejournals.org

Hybrid-Electric, Heavy-Fuel Propulsion System for Small Unmanned


Aircraft
Kyle Merical, Troy Beechner, and Paul Yelvington
Mainstream Engineering Corp.

ABSTRACT
A series hybrid-electric propulsion system has been designed for small rapid-response unmanned aircraft systems (UAS)
with the goals of improving endurance, providing flexible and responsive electric propulsion, and enabling heavy fuel
usage. The series hybrid architecture used a motor-driven propeller powered by a battery bank, which was recharged by
an engine-driven generator, similar to other range-extended electric vehicles. The engine design focused on a custom,
two-stroke, lean-burn, compression-ignition (CI), heavy-fuel engine, which was coupled with an integrated starter alternator
(ISA) to provide electrical power. The heavy-fuel CI engine was designed for high power density, improved fuel efficiency,
and compatibility with heavy fuels (e.g., diesel, JP-5, JP-8). Commercially available gasoline spark-ignition engines and
heavy-fuel spark-ignition engines were also considered in the trade study. The series hybrid configuration allowed the
engine to be mechanically decoupled from propeller, so that the engine could be operated at the load/speed condition for
peak fuel-conversion efficiency. An energy-dense rechargeable battery pack was used to store energy and allow the UAS
to operate with the engine shut off, which provided an engine-off operating mode. The ISA allowed re-starting the engine in
flight without the need for a separate starter motor.

Simulation-based design tools were developed, and trade studies were performed for the various system components. The
series hybrid UAS outfitted with the custom diesel engine demonstrated endurance improvements, due to additional
benefits netted from the improved engine efficiency. Development of the hybrid propulsion system is ongoing, with current
efforts focused on reducing system mass, packaging, and gearing up for a future hardware demonstration.

CITATION: Merical, K., Beechner, T., and Yelvington, P., "Hybrid-Electric, Heavy-Fuel Propulsion System for Small Unmanned
Aircraft," SAE Int. J. Aerosp. 7(1):2014, doi:10.4271/2014-01-2222.

INTRODUCTION highly efficient electric machine can be used to complement


the ICE and reduce the time the ICE spends at inefficient
The goal of this ongoing research program is to develop a
operating conditions, thus conserving fuel.
small hybrid-electric rapid-response UAS that is capable of
cruising for extended periods and performing “fast dash”
The design of a series hybrid electric propulsion system for a
maneuvers. Additionally, the UAS is being designed for heavy
small UAS is presented. The propulsion system uses a
fuels (primarily JP-5/8 jet fuels) rather than gasoline or other
two-stroke, compression-ignition, heavy-fuel engine connected
fuels, to simplify the supply chain for military operations.
to an ISA, to generate electrical power. The generated power is
Commercial applications are also anticipated, for example, to
then used to either recharge a lithium-ion battery pack, or
support oil exploration or marine wildlife observation.
directly power a brushless DC (BLDC) motor that is connected
to the pusher propeller. This type of propulsion system allows
Hybrid propulsion systems exploit the best attributes of multiple
the engine operation to be decoupled from the propeller, thus
types of power sources to more efficiently propel a vehicle.
allowing the engine to function as a generator, running at its
Common hybrid power sources include internal combustion
most efficient condition. The hybrid architecture also allows the
engines (ICEs), electric motors, hydraulic pumps, and
UAS to be operated with the engine turned off.
flywheels. Hybrid electric vehicles (HEVs) that combine energy
dense internal combustion engines and highly efficient electric
A true compression-ignition two-stroke heavy-fuel engine was
machines are among the most common hybrids. The HEV
designed for the subject application using one-dimensional
offers more utility than a pure electric vehicle (EV) because the
engine cycle simulation software (Ricardo WAVE). This custom
energy dense hydrocarbon fuel aboard the HEV allows for an
engine was designed to replace the modified hobby aircraft
extended range and can be quickly refueled. The HEV also
engines that are currently used in small UASs (nominally 30-60
improves upon a conventional propulsion system because a
126
Downloaded from SAE International by Univ of California, Friday, July 27, 2018

Merical et al / SAE Int. J. Aerosp. / Volume 7, Issue 1 (September 2014) 127

lb [14-27 kg] gross take-off weight, GTOW). The power density 30 lb UAS using a parallel hybrid-electric propulsion system
of the two-stroke cycle is ideal for aviation. Direct fuel injection was detailed in Harmon, Frank, and Chattot [2]. This vehicle
is used to prevent short circuiting of fuel out the exhaust of the used a downsized four-stroke spark-ignition (SI) ICE and
engine, as this leads to excessive fuel consumption with electric motor coupled to the propeller via a belt and pulley
carbureted two-stroke engines. The compression-ignition (i.e., system. A one-way bearing on the ICE pulley allowed the
diesel) cycle also increases thermal efficiency and allows for vehicle to operate in electric-only mode, but a separate starter
practical use of heavy fuels, such as JP-8, thus adhering to the was required to restart the engine. Simulations predicted that
single fuel forward concept of the U.S. military. A high efficiency energy consumption was reduced by 54% for a one hour
ISA was used to both start the engine, as well as generate mission and 22% for a three hour mission when compared to
electricity, after the engine is firing. Combining the starter and the baseline UAS. However, the low power density of this
alternator into a single component reduces space, weight, and aircraft limited its predicted top speed to 75 mi/h (120 ki/h).
parts count. This design was refined and bench tested by Hiserote and
Harmon [3] and Greiser et al. [4]. The complete propulsion
system was retrofitted into a small aircraft and underwent taxi
BACKGROUND tests by Ausserer and Harmon [5]. Flight tests were pending as
Three potential HEV architectures are shown in Figure 1. of 2012.
Parallel HEVs use an ICE and electric motor simultaneously
connected to a drivetrain to deliver power to the propeller. Most Design of a parallel hybrid propulsion system for the AAI
of the previous design studies of hybrid aircraft have Aerosonde UAS was presented by Hung and Gonzalez [1].
considered this configuration. The portion of the required This propulsion system consisted of a small four-stroke
propeller power that comes from the ICE versus the electric spark-ignition engine and electric motor connected to a
motor is varied at different operating points to maximize continuously variable transmission (CVT). A rule-based ideal
efficiency. The series HEV architecture uses the ICE solely to operating line strategy was used to achieve a predicted 6.5%
generate electricity, which is then used to power the electric reduction in fuel consumption during a nominal mission,
drive motor that is directly connected to the propeller. compared to a vehicle with conventional propulsion system.

The more complex power-split hybrid is a type of parallel hybrid The series hybrid architecture has already been used in the
architecture that uses a planetary gear set to split power Diamond Aircraft DA36 E-Star manned aircraft, as well as
among an ICE, generator/motor, and electric drive motor. This automobiles such as the Chevrolet Volt. The current Diamond
architecture allows for the most efficient use of all devices but Aircraft DA36 E-Star 2 is powered by a propeller attached to a
requires additional hardware and more advanced control. 65 kW electric motor. Electrical power is supplied by a 30 kW
There are currently no prototype hybrid UASs that use the Austro AE50R rotary engine and generator, yielding a 25%
power split architecture. reduction in fuel consumption over a conventional aircraft.

Many different types of control have been used to minimize


energy consumption of HEV powertrains. Simple rule-based
control is viable but potentially difficult to implement because,
as more scenarios are considered, the number of necessary
rules becomes cumbersome. Fuzzy logic controllers still use
rule-based control but use continuous output values between 0
and 1 (instead of a discrete output of 0 or 1). Fuzzy logic
control is robust but can be complicated for systems with many
input variables. Neural network controllers use optimization
algorithms to assign relative importance between controller
inputs. The selected controller response is the most optimal
solution at that time. The equivalent consumption minimization
Figure 1. Series, parallel, and power-split HEV configurations, adapted strategy (ECMS) minimizes a cost function describing the
from Hung and Gonzalez [1]. equivalent fuel consumption of a parallel HEV.

The relatively low specific energy of battery packs has Rule-based control was chosen for this effort because the
traditionally prevented HEVs from being used in aviation. control of a series hybrid vehicle is much simpler than control
However, various hybrid UAS propulsion architectures and of a parallel or power-split hybrid. In a series hybrid, the engine
applications have been simulated [1, 2, 3, 4, 5, 6, 7, 8, 9, 10, is either turned on or off to generate electrical power. There is
11, 12, 13, 14, 15, 16, 17, 18, 19, 20]; although, to our no power split between the engine and electric motor to
knowledge, none of these have yet been flown. The design of a determine, so complex control methods are not necessary.
Downloaded from SAE International by Univ of California, Friday, July 27, 2018

128 Merical et al / SAE Int. J. Aerosp. / Volume 7, Issue 1 (September 2014)

ENERGY ANALYSIS AND COMPONENT 95% energy conversion efficiency for the ISA and 97%
SIZING efficiency for power conditioning) to the ISA while the batteries
are simultaneously discharged at a 2C rate. Assuming that the
Initial steady-state calculations were performed to estimate the engine achieves a brake thermal efficiency of 20% at 4500
power requirements for cruise and dash conditions, which RPM and an air-fuel-ratio (AFR) of 25, an engine displacing 40
provided initial constraints on the engine size, and battery pack cm3 is required. Below approximately 55 mi/h (88.5 km/h), the
capacity. electrical power requirement can be met with the batteries
alone based on the prescribed 2C maximum discharge rate.
Power Requirements
To explore the difference in performance when using
Aerodynamic power requirements were calculated for the
compression ignition engines of differing sizes, engines
hybrid UAS during steady level flight, at varying vehicle
displacing 28 cm3, 54 cm3, and 80 cm3 were also modeled. All
speeds, assuming a typical fixed-wing airframe and fixed-pitch
four diesel engines were modeled in Ricardo WAVE, along with
propeller. It was assumed that the hybrid UAS had a GTOW of
a commercially available, 28 cm3, two-stroke, gasoline spark-
55.8 lb (25.3 kg), including fuel and payload, for all cases
ignition engine and a heavy-fuel variant of the same spark-
considered. An appropriately sized commercial propeller was
ignition engine. The WAVE models were exercised over the
selected to match the thrust requirements and motor speed for
load and speed range of each engine, while also varying
the hybrid UAS.
altitude. The data was imported into MATLAB and a script
determined the engine operating point with the best brake-
A notional drag polar and propeller efficiency curves were used
specific fuel consumption (BSFC), as well as the operating
to estimate the aerodynamic power. The aerodynamic power
point with the highest peak power. Table 1 summarizes the
divided by the energy conversion efficiencies of the propeller,
predicted performance of each engine in both generator mode
drive motor, and motor-drive power electronics, yielded the
and peak power mode.
electrical power requirement of the drivetrain.

The masses of the commercial engines were provided by the


Battery Pack Sizing manufacturer. A trade study was carried out to estimate the
The battery pack was the sole energy source for the vehicle mass of each custom-designed engine. Figure 2 shows a
during engine-off mode, so this requirement helped determine scatter plot of engine mass as a function of engine
the appropriate size for the battery pack. The goal was to displacement for various commercial small engines. Due to
specify a battery pack using commercially available battery limited availability, data was only accessible for three two-
cells that would allow the UAS to remain in engine-off mode stroke, heavy-fuel, single-cylinder engines. Of these three
while at the target cruise speed for one hour. engines, one was a true compression-ignition (CI) engine
(Cosworth), while the other two were spark-ignition (SI)
We estimated that 596 W are consumed by the electric engines. Figure 2 shows that gasoline two-stroke engines are
drivetrain during cruise, so 596 Wh of battery capacity are lighter than heavy-fuel engines of the same displacement
required for the UAS to cruise in silent mode for one hour. The because of their relative simplicity. Using heavy fuels requires
desired pack voltage was 60 V, in order to minimize power more complex fuel delivery systems as well as fuel heating
transmission losses and negate the need to boost the voltage elements. However, the higher efficiency of the two-stroke CI
supplied to the electric drive motor. engine can offset the weight penalty in a UAS application. The
additional weight can be minimized by using a downsized CI
A commercial Panasonic lithium-ion cell was selected as the engine or by carrying less fuel, as less is required for the same
building block for the battery pack because of its high energy flight duration.
density (243 Wh/kg). This cell nominally provides 3.6 V and
has 3.35 Ah of capacity. To closely match the target 60 V pack Table 1. Predicted Performance of Modeled Engines
voltage, a string of 16 cells in series was used. The energy
requirement of 596 Wh was divided by the pack voltage of 57.6
V to yield a total pack capacity requirement of 10.3 Ah. To
nearly match this target, three parallel strings were required
giving a total capacity of 10.05 Ah.

Initial Engine Sizing


Engine power demands determined the ideal size of the
next-generation, purpose-built diesel two-stroke UAS engine.
To offset the additional weight of the hybrid propulsion system
and achieve the target top speed, the propulsion system must
supply 1884 W of electrical power to the motor. The engine
must therefore supply 1432 W of shaft power (considering a
Downloaded from SAE International by Univ of California, Friday, July 27, 2018

Merical et al / SAE Int. J. Aerosp. / Volume 7, Issue 1 (September 2014) 129

Compression-Ignition Two-Stroke Engine


A CI (i.e., diesel cycle) engine suitable for use in the hybrid
UAS propulsion system was designed. The two-stroke cycle
was selected because of the increased power density
compared to the four-stroke cycle. However, because SI
two-stroke engines achieve poor fuel economy, compression-
ignition and direct-injection were employed. The CI approach
takes advantage of a higher allowable compression ratio and
corresponding greater thermal efficiency. Additionally, because
the fuel is directly injected into the cylinder after all ports are
closed, no fuel can “short circuit” during the gas exchange
process and exit the exhaust port without being burned. Finally,
the CI engine is much better suited for the low octane number
of the JP-8 fuel (ON∼15) compared to an SI engine. The major
challenge facing small-displacement CI engines is poor fuel
Figure 2. Engine mass versus displacement for commercially available
mixture preparation due to impingement of the fuel spray on
heavy-fuel and gasoline engines.
the piston or cylinder walls. Complementary efforts to develop
The linear fit to the relationship between engine mass and fuel injectors for small UAS engines are also underway.
displacement, shown in Figure 2, was evaluated for each
custom-designed heavy-fuel engine. Table 2 shows the After the basic architecture of the engine was selected, the
predicted mass of each of these CI engines. These initial initial calculations described previously were used to specify
predictions do not explicitly consider the additional mass the dimensions of the 28 cm3, 40 cm3, and 80 cm3 engines.
required to strengthen components to handle the higher peak These dimensions were used to create 1-D models of the
pressures in the CI engine, although this effect is somewhat engine using WAVE. The engine models were used to evaluate
captured by including the Cosworth CI engine in the the performance of each engine across its entire operating
correlation. The difference in mass between each engine and range of engine speed and load. Figure 3 shows the BSFC of
the baseline gasoline engine is also shown. The additional the 28 cm3 engine at sea level. The peak brake thermal
mass of each engine was then subtracted from the initial fuel efficiency of this engine was over 28%.
mass carried on the baseline UAS to maintain the same
GTOW. This means that the UASs equipped with larger The benefit of using an engine with a displacement that is
engines carried substantially less fuel. larger than necessary to match the power demands of the UAS
during cruise is evident from Figure 3. The BSFC does not
Table 2. Predicted Engine Mass. change significantly as air-fuel ratio (AFR) varies at moderate
engine speeds (∼3500 RPM). This allows for an engine that
can turn down well and function efficiently at both low and high
power levels. Ideally, during cruise the generator should be
operated at the most efficient operating point because the
engine spends the majority of the time in this mode.

COMPONENT DESIGN AND MODEL


FORMULATION
Dynamic models of the hybrid UAS components were
constructed and used to construct a system-level model. The
system-level model was used to estimate the hybrid UAS
performance for various flight profiles.
Figure 3. BSFC (kg/kWh) map for 28 cm3 CI engine at sea level
Downloaded from SAE International by Univ of California, Friday, July 27, 2018

130 Merical et al / SAE Int. J. Aerosp. / Volume 7, Issue 1 (September 2014)

Before the CI engines were modeled, a model of the heavy-fuel shaft. This position dependent toque is shown in Figure 4
variant of the commercial SI engine was created in WAVE, and where the 0.2 second timescale corresponds to 750 degrees of
the model was calibrated using experimental data from the rotation for the commercial gasoline engine.
manufacturer. The experimental conditions were simulated
using the WAVE model of the commercial heavy-fuel engine.
Model parameters such as valve discharge coefficients were
then calibrated to improve model agreement. After the
commercial heavy-fuel engine model was calibrated, this same
engine was used as the basis for creating the WAVE models
for the CI engines previously discussed.

Integrated Starter Alternator


The ISA was designed to suit the commercial gasoline engine,
so that the hybrid propulsion system may be built independent Figure 4. Piecewise linear approximation of the required starting torque
of the custom-designed diesel engines. The assumed for the commercial engine over 750 degrees of rotation.
alternator requirements for the hybrid UAV are shown in Table
3. During normal operation, the ISA will function as a power It can be seen from Figure 4 that the commercial gasoline
source to either recharge the system's Li-ion batteries or engine places approximately 6 Nm of torque onto the alternator
provide power to the propeller motor. The second function of shaft during the compression stroke. Using a two-stroke engine
the ISA is to provide the required starting torque for the with a compression stroke every crankshaft revolution required
commercial engine. The selected motor type is a brushless the system control ramp-rate to be quite rapid. However,
three-phase permanent magnetic (PM) motor that interfaces starting with the ISA was viable because the combined inertia
with a bi-directional three phase active rectifier. of the engine and optimized alternator was small.

Table 3. Basic ISA Design Specifications. In order to investigate the performance of the starting
controller, the system model was run for 0.2 seconds. Figure 5
shows the engine speed during startup. The slow-down of the
engine during the compression strokes is noted on the figure.
The transient analysis demonstrated the ability of the self-
starting controller to overcome the peak torques of the initial
compression stroke allowing the engine to start.

The alternator design approach sought to minimize the overall


volume and weight while still producing the required starting
torque. The specifications of the final ISA design are shown in
Table 4.

Table 4. Specifications of UAV PM Alternator.

Figure 5. Engine speed during starting with the ISA.

Battery Pack
A dynamic model (in Simulink) was created to verify starting The electrical current integration method was used to estimate
operation of the UAV hybrid system, including the custom battery state of charge (SoC), as described by (1) where i is
designed PM alternator and commercial engine. The model the current and Qbatt is the battery capacity. The battery pack
was adapted from previously validated models which we have voltage was modeled using a grey-box approach, where
used to design motors and alternators in the past. additional terms were added to physics-based equations
describing the general behavior of a lithium-ion battery pack, in
In order to decrease simulation times, the engine model was order to obtain a better fit between experimental data and
simplified such that its output is a piecewise linear crank- model outputs. The battery voltage was determined using (2),
position dependent torque, which was applied to the alternator where E0 is the nominal battery voltage and Tbatt is the battery
Downloaded from SAE International by Univ of California, Friday, July 27, 2018

Merical et al / SAE Int. J. Aerosp. / Volume 7, Issue 1 (September 2014) 131

temperature. Battery temperature was obtained by performing all flight conditions are assumed to be at constant vehicle
an energy balance on the battery, considering heat produced speed, and therefore modeling the transient behavior of the
by the battery, and heat transferred to the surroundings. UAS was not a primary concern.

A charge management system (CMS) was also created, and


applied rule-based control to manage the battery pack and
(1) engine operation. Battery pack charging and discharging and
engine operation were managed according to power demands,
battery pack SoC and voltage, and whether or not the engine-
(2) off mode had been requested.

The calibrated battery model produced model results that


closely matched manufacturer's experimental data. Figure 6
shows a comparison of experimental data and model outputs,
for varying battery cell discharge rates.

Figure 7. Complete system model block diagram.

Simulation Results
The maximum flight endurance of each UAS configuration was
simulated. The simulation was first completed with the UAS
equipped with the conventional, non-hybrid, propulsion system
and each commercial engine. The maximum endurance of the
Figure 6. Single-cell battery model discharge curves compared to UAS with the hybrid propulsion system was then simulated with
experimental manufacturer's data. each commercial engine, as well as the four custom-designed
diesel engines. This paper focuses on the hybrid
Propeller Drive Motor configurations-Table 5 shows the configuration of each hybrid
UAS propulsion system tested.
A brushless DC (BLDC) drive motor was specified that could
produce the torque and speed required to drive the propeller of
Table 5. UAS Hybrid Propulsion Systems Simulated.
the UAS. At the desired top speed, the propeller and electric
motor needed to rotate at 7900 RPM and required 2381 W of
electrical power. Given these parameters, a commercially
available BLDC motor was identified for use on the hybrid
propulsion system.

System-Level Modeling
Each component model previously described was implemented
into a system-level model of the entire UAS, as shown in
Figure 7. A flight cycle submodel that receives the given flight
profile, specified by a vehicle speed, climb speed, and binary
Each UAS began the simulation with a full fuel tank and fully
“engine-off” mode request flag, was also added to the system
charged battery pack. The UAS was flown at a constant cruise
model. The static backward-looking system model works
speed (55 mi/h) at sea level until the fuel supply and battery
backward from the given flight conditions to determine first the
pack energy were both completely exhausted. Figure 8 shows
aerodynamic power required, and then the electrical power
the results of this test for a hybrid UAS equipped with the 28
required, along with all other applicable parameters. The
cm3 commercial gasoline engine. The charge management
backward-looking model is well suited to this approach
system (CMS) regulated the battery pack state-of-charge
because it does not introduce the tracking error associated with
(SoC) as normal until the fuel was depleted and then allowed
a forward-looking model that manipulates the “pilot” controls to
the battery to be fully discharged and stopped the simulation.
attempt to meet the desired flight profile. Also, the majority of
Downloaded from SAE International by Univ of California, Friday, July 27, 2018

132 Merical et al / SAE Int. J. Aerosp. / Volume 7, Issue 1 (September 2014)

The endurance and fuel consumption for the six cases are
plotted in Figure 9. The custom CI engines are shown with
circle markers, and the commercial SI engines are shown with
triangle markers. The markers are colored by the initial fuel
mass (i.e., fuel capacity). The fuel consumption for the CI
engines was very similar due to the similarity of the brake
thermal efficiency in generator mode for these cases, which
only varies over the narrow range from 27.3% to 28.7%. This
similarity is due to the fact that the most efficient operating
point was selected for each case subject to a minimum
threshold power for acceptable charging times in generator
mode. The large difference in fuel consumption between CI
and SI cases is evident from the figure. The best endurance
was observed for the 28 cm3 CI engine (Case 3) due to its low
fuel consumption and light weight, which allowed for a higher
fuel capacity. This analysis highlights the unmistakable impact
Figure 8. UAS maximum flight endurance test. of fuel capacity on endurance, regardless of the propulsion
technology used.
Table 6 shows the predicted fuel consumption at cruise and
maximum flight endurance of the hybrid UAS when equipped
with each engine. The fuel consumption is an average over the
entire flight, which includes periods when the engine is both on
and off. As shown in Table 6, the endurance for a hybrid UAS
with the commercial, gasoline SI engine (Case 1) was 15.1 hr.
The endurance for a hybrid UAS with the commercial, heavy-
fuel SI engine (Case 2) was 12.3 hr, 18.5% lower than the
gasoline engine. The reduced endurance was due to the lower
brake thermal efficiency of the heavy-fuel SI engine compared
to the gasoline SI engine. The fuel consumption for the CI
engines (Cases 3-6) was about 50% lower than the
commercial gasoline SI engine. The endurance was also
improved with the custom CI engines compared to the
commercial SI engine even though those engines carried less
fuel compared to the gasoline SI engine to compensate for the
increased weight of the engine. Although not considered in this
trade study, the 80 cm3 CI engine produced more power than
the baseline commercial SI engine (2.7 kW vs. 2.25 kW) and Figure 9. Trade study of UAS propulsion systems. Commercial engines
therefore could have potentially supported a somewhat higher shown as triangles; prototype CI engine designs shown as circles.
GTOW allowing for a larger fuel capacity. Although the 80 cm3
engine had a considerably larger displacement, the maximum The top speeds for the six cases were predicted to vary by less
operating speed was 4500 rpm (typical maximum for diesels) than 10% even though the peak power of the engines varied
compared to 8500 rpm for the baseline SI engine, which limited by almost a factor of three. The required aerodynamic power
power output (see Table 1). increased sharply with the square of vehicle speed due to drag,
resulting in minimal change in top speed with engine power. An
Table 6. Predicted Performance of the Hybrid UAS with Different improved airframe with better high-speed aerodynamics would
Engines. have shown a larger variance in top speed with engine power.
Regardless, the top speed for the hybrid UAS with the 80 cm3
custom CI engine exceeded the top speed of the commercial
engines due to its higher peak power (see Table 1).

ISA HARDWARE DEMONSTRATION


The ability to start the engine with the ISA was identified as a
high risk element in the hybrid propulsion system. Therefore a
hardware demonstration of this functionality was undertaken.
An existing 2 kW diesel generator with a flywheel-ISA
developed previously at Mainstream Engineering was used as
the demonstration platform in lieu of the target UAS engine/ISA
Downloaded from SAE International by Univ of California, Friday, July 27, 2018

Merical et al / SAE Int. J. Aerosp. / Volume 7, Issue 1 (September 2014) 133

hardware, which has not yet been fabricated. A high-level


schematic of the experiment is shown in Figure 10 and a
picture of the hardware is shown in Figure 11. The control
strategy was first tested in software and then downloaded to a
custom control board, which includes a 60 MHz digital signal
processor (DSP). This DSP ran the self-starting and rectifier
control algorithms, interfaced with system sensors, and in
general controlled operation of the engine/generator system. In
order to start the engine, the controller commands torque
producing current (3-phase) be injected into the alternator, thus
spooling up the engine. The engine speed and crankshaft
position were monitored using an optical encoder.

Figure 12. Phase B current (blue) and alternator back EMF voltage
(yellow) during engine starting.

Figure 10. High-level experimental system block diagram. SUMMARY


The series hybrid-electric propulsion system using a custom-
designed 28 cm3 compression-ignition (CI) engine was
predicted to give an endurance of 23.7 hours at cruise speed.
By comparison, a hybrid UAS using a commercial gasoline SI
engine was predicted to have an endurance of 15.1 hours.
While utilizing a more efficient diesel-cycle CI engine provided
a larger benefit for endurance than the commercial gasoline SI
engine, development of small CI engines poses technical
challenges related to downsizing of the high-pressure fuel
injector. In general, hybrid UAS offer a number of other benefits
not obtainable with conventional propulsion (diesel or
otherwise), such as providing an engine-off operating mode,
improving the transient response of the aircraft, and improving
the durability of the engine by cycling it off.

The use of an integrated starter-alternator appears to be a


Figure 11. Single-cylinder diesel engine, ISA, and power electronics viable method for engine starting and power generation in a
used as the demonstration platform. hybrid UAS thereby eliminating the need for a separate starter
motor. Future analyses will focus on minimizing the system
Figure 12 shows test data that was collected while successfully
weight and determining system performance during nominal
starting the engine with the ISA starting algorithm. The data
flight profiles, which will include take off, climb out, dash,
shows the Phase B current (blue) used to crank the engine,
cruise, and descent sequences.
which goes to zero once the engine starts and reaches the
crossover speed threshold. The alternator back EMF voltage is
also shown (yellow), which highlights the fact that once the REFERENCES
engine starting algorithm shuts down (and commands zero 1. Hung, J.Y. and Gonzalez L.F., “On parallel hybrid-electric
current), the engine continues to increase its speed, resulting propulsion system for unmanned aerial vehicles,” Progress
in Aerospace Sciences, 2012, 51:1-17, doi:10.1016/j.
in a back EMF voltage of increasing magnitude and frequency. paerosci.2011.12.001.
After a few more seconds, Figure 12 shows that the engine has 2. Harmon, F.G., Frank A.A., and Chattot J.-J., “Conceptual design
started and reached its steady state operating speed of 3600 and simulation of a small hybrid-electric unmanned aerial vehicle,”
Journal of Aircraft, 2006, 43(5): 1490-1498, doi:10.2514/1.15816.
RPM. This demonstration showed that our ISA starting
3. Hiserote, R.M. and Harmon F.G., “Analysis of hybrid-electric
algorithm was viable. Also, the observed current required to propulsion system designs for small unmanned aircraft systems,”
start the demonstration engine agreed well with the simulation 8th Annual International Energy Conversion Engineering
Conference, 2010, Nashville, TN, USA.
predictions, thereby validating the dynamic model used to size
4. Greiser, C.M., Mengistu I.H., Rotramel T.A., and Harmon F.G.,
the alternator and power electronics for the hybrid UAS. “Testing of a parallel hybrid-electric propulsion system for use in
a small remotely-piloted aircraft,” 9th Annual International Energy
Conversion Engineering Conference, IECEC 2011, 2011, San
Diego, CA, USA.
Downloaded from SAE International by Univ of California, Friday, July 27, 2018

134 Merical et al / SAE Int. J. Aerosp. / Volume 7, Issue 1 (September 2014)

5. Ausserer, J.K. and Harmon F.G., “Integration, validation, and 15. Li, Y., Liu L., Ma X., and Tu H., “Design of hybrid electric
testing of a hybrid-electric propulsion system for a small remotely- propulsion system for long endurance small UAV,” 10th Annual
piloted aircraft,” 10th Annual International Energy Conversion International Energy Conversion Engineering Conference, IECEC
Engineering Conference, IECEC 2012, 2012, Atlanta, GA, USA. 2012, 2012, Atlanta, GA, USA.
6. Capata, R., Marino L., and Sciubba E., “A hybrid propulsion 16. Lieh, J., Behbahani A., and Hoying J., “Modeling of hybrid
system for a high-endurance UAV: Configuration selection electric UAV propulsion system in simulink,” Proceedings of the
and aerodynamic study,” ASME 2011 International Mechanical International Instrumentation Symposium, 2012, San Diego, CA,
Engineering Congress and Exposition, IMECE 2011, 2011, USA.
Denver, CO, USA. 17. Lieh, J., Spahr E., Behbahani A., and Hoying J., “Design of hybrid
7. Earon, E., Rabbath C.A., and Apkarian J., “Pushing the envelope: propulsion systems for unmanned aerial vehicles,” 47th AIAA/
A novel hybrid vehicle design and real-time control concept,” ASME/SAE/ASEE Joint Propulsion Conference and Exhibit 2011,
Proceedings of the 9th IASTED International Conference on 2011, San Diego, CA, USA.
Control and Applications, 2007, Montreal, QC, Canada. 18. Schoemann, J. and Hornung M., “Modeling of hybrid-electric
8. Harmon, F.G., Frank A.A., and Joshi S.S., “Application of propulsion systems for small unmanned aerial vehicles,” 12th
a CMAC neural network to the control of a parallel hybrid- AIAA Aviation Technology, Integration and Operations (ATIO)
electric propulsion system for a small unmanned aerial Conference, 2012, Indianapolis, IN, USA.
vehicle,” Proceedings of the International Joint Conference on 19. Siemens. Le Bourget: Electric Hybrid Drives for Aircraft. 2013
Neural Networks, 2005, Montreal, QC, Canada, doi:10.1109/ [accessed 2013]; Available from: http://www.siemens.com/
ijcnn.2005.1555856. innovation/en/news/2013/e_inno_1318_1.htm.
9. Harmon, F.G., Frank A.A., and Joshi S.S., “The control of a 20. Verstraete, D., Cazzato L., and Romeo G., “Preliminary design
parallel hybrid-electric propulsion system for a small unmanned of a fuel-cell-based hybrid-electrical UAV,” 28th Congress of the
aerial vehicle using a CMAC neural network,” Neural Networks, International Council of the Aeronautical Sciences 2012, ICAS
2005, doi:10.1016/j.neunet.2005.06.030. 2012, 2012, Brisbane, Australia.
10. Harmon, M.F.G., Frank A.A., and Chattot J.J., “Parallel hybrid-
electric propulsion system for an unmanned aerial vehicle,”
AUVSI's Unmanned Systems North America 2004 - Proceedings, CONTACT INFORMATION
2004, Anaheim, CA, USA. For more information, contact the authors at
11. Hrad, P.M. and Harmon F.G., “Conceptual design tool for Micro
Air Vehicles with hybrid power systems,” 8th Annual International
Energy Conversion Engineering Conference, 2010, Nashville, TN, Mainstream Engineering Corporation
USA. Rockledge, FL, USA
12. Karunarathne, L., Economou J.T., and Knowles K., “Fuzzy
logic control strategy for Fuel Cell/Battery aerospace phone: 321-631-3550
propulsion system,” 2008 IEEE Vehicle Power and Propulsion pyelvington@mainstream-engr.com
Conference, VPPC 2008, 2008, Harbin, China, doi:10.1109/
vppc.2008.4677772.
13. Karunarathne, L., Economou J.T., and Knowles K., “Model ACKNOWLEDGMENTS
based power and energy management system for pem fuel cell/
Li-Ion battery driven propulsion system,” IET Conference, 2010, Support for this effort was provided by the U.S. Air Force under
Brighton, United Kingdom, doi:10.1049/cp.2010.0088. Contract No. FA8650-13-M-2377. The authors thank Thomas
14. Koster, J.N., Balaban S., Hillery D.R., Serani E., Velazco A.,
Humbargar C., Goodman C., Brewer A., Johnson M., Kosyan M., Barnett, Ryan Miller, and Michael Rottmayer at the Air Force
Nasso D., Price J., Wiley T., Zhao R., Munz C.-D., Kraemer E., Research Laboratory for helpful discussions and information
Kurz H., Arenz M., Pfeifer D., Seitz M., Wong K.C., Verstraete on small unmanned aircraft.
D., and Lehmkuehler K., “Rapid, international design and test of
a hybrid-powered, blended-wing body unmanned aerial vehicle,”
11th AIAA Aviation Technology, Integration, and Operations (ATIO)
Conference, 2011, Virginia Beach, VA, USA, doi:10.2514/6.2011-
6964.

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 International.

Positions and opinions advanced in this paper are those of the author(s) and not necessarily those of SAE International. The author is solely responsible for the content of the
paper.

You might also like