79587128
79587128
79587128
by
Thomas Strele
December 2016
ABSTRACT
As electric powered unmanned aerial vehicles enter a new age of commercial viability,
market opportunities in the small UAV sector are expanding. Extending UAV flight time through a
combination of fuel cell and battery technologies enhance the scope of potential applications. A
brief survey of UAV history provides context and examples of modern day UAVs powered by fuel
cells are given. Conventional hybrid power system management employs DC-to-DC converters to
control the power split between battery and fuel cell. In this study, a transistor replaces the DC-to-
DC converter which lowers weight and cost. Simulation models of a lithium ion battery and a
proton exchange membrane fuel cell are developed and integrated into a UAV power system
model. Flight simulations demonstrate the operation of the transistor-based power management
scheme and quantify the amount of hydrogen consumed by a 5.5 kg fixed wing UAV during a six
hour flight. Battery power assists the fuel cell during high throttle periods but may also augment
fuel cell power during cruise flight. Simulations demonstrate a 60 liter reduction in hydrogen
consumption when battery power assists the fuel cell during cruise flight. Over the full duration of
the flight, averaged efficiency of the power system exceeds 98%. For scenarios where inflight
battery recharge is desirable, a constant current battery charger is integrated into the UAV power
system. Simulation of inflight battery recharge is performed. Design of UAV hybrid power systems
must consider power system weight against potential flight time. Data from the flight simulations
are used to identify a simple formula that predicts flight time as a function of energy stored
onboard the modeled UAV. A small selection of commercially available batteries, fuel cells, and
compressed air storage tanks are listed to characterize the weight of possible systems. The
formula is then used in conjunction with the weight data to generate a graph of power system
weight versus potential flight times. Combinations of the listed batteries, fuel cells, and storage
tanks are plotted on the graph to evaluate various hybrid power system configurations.
i
DEDICATION
ii
ACKNOWLEDGMENTS
I would like to thank Dr. Changho Nam for his guidance and support during this project.
iii
TABLE OF CONTENTS
Page
CHAPTER
5.3 Conclusion…………………….……………………………………………….114
APPENDIX
O OBSERVER_1 DIAGRAM……………………………………………………………………157
P OBSERVER_2 DIAGRAM……………………………………………………………………..159
AA MATLAB SCRIPT TO PLOT FLIGHT TIME VERSUS POWER SYSTEM WEIGHT ........ 192
vi
LIST OF TABLES
Table Page
7. Hydrogen Consumed and Battery Final SOC for Different Battery Capacities .............. 93
9. Comparison of Simulation Data and Flight Time Approximation Formula .................... 101
12. Ninja Pro v2 Tank Weights and Storage Capacities .................................................... 103
vii
LIST OF FIGURES
Figure Page
7. ASU Composite UAV CAD Assembly (left), ASU Composite UAV Disassembly (right) 21
9. Battery Type Volumetric Energy Density versus Specific Energy Density ..................... 23
26. Percent Error of Lee Model for Different Hydrogen Pressures ...................................... 64
viii
Figure Page
27. MEA Resistance versus Current Density for Different Relative Humidity Values ........ 66
32. PEMFC Dynamic Response for Range of Double Layer Capacitance ......................... 70
45. Fuel Cell and Battery Voltage and Current versus Time (hybrid = 0 and 1.5) .............. 88
49. Battery SOC and Hydrogen Consumed for Different Battery Capacities ...................... 93
51. Hybrid Power System Performance for 11 Ah Battery Expanded View ........................ 95
54. Fuel Cell and Battery Voltage and Current during Inflight Recharge ............................ 98
55. Battery SOC and Current during Inflight Battery Recharge ........................................... 98
57. Flight Time versus Weight for Power System Configurations ..................................... 105
64. Oscilloscope Capture of ESC Voltage and Current Motor Signals (zoomed in) …….183
x
LIST OF SYMBOLS AND ACRONYMS
Symbol Page
INTRODUCTION
Unmanned Aerial Vehicles (UAV) may be loosely described as an aircraft without a pilot
onboard. A UAV may be piloted remotely via radio control or fly autonomously using an onboard
autopilot system. In some cases, both forms of flight control are present. That is to say radio
control is present in addition to an onboard autopilot system. Refining the definition of a UAV, a
propulsion system must also be present onboard the aircraft (“DOD dictionary”, 2007). A
propulsion system may consist of a motor with a propeller or a jet engine. This second
requirement excludes flying projectiles such as bullets and artillery shells. Typically, the term UAV
implies a powered unmanned aircraft, either fixed wing or rotary wing. According to the U.S.
A powered, aerial vehicle that does not carry a human operator, uses aerodynamic forces
to provide vehicle lift, can fly autonomously or be piloted remotely, can be expendable or
recoverable, and can carry a lethal or nonlethal payload. Ballistic or semi-ballistic
vehicles, cruise missiles, and artillery projectiles are not considered unmanned aerial
vehicles. (“DOD dictionary”, 2007)
Note that all ballistic or semi-ballistic rockets and missiles are excluded from the UAV definition.
For the special case of cruise missiles (non-ballistic), where aerodynamic forces provide
significant lift, the DOD definition explicitly excludes them from the UAV definition. There are
many examples of both fixed wing and rotary wing UAVs, but this paper will focus only on fixed
wing UAVs.
UAVs have been around since before the first manned powered airplane flight of the
Wright brothers in 1903 (“Wright Brothers”, n.d.). The first recorded UAV flight took place in 1896
when Samuel Langley launched an unmanned, steam powered aircraft from a house boat on the
Potomac River. The aircraft, named the “Aerodrome No. 5”, flew for about 90 seconds and
covered a distance of 1005 m (“Langley Aerodrome Number 6”, n.d.). The Aerodrome weighed
11.4 kg and had a wingspan of 4.2 m. The engine consisted of a single cylinder steam engine
Navy and U.S. Army funded UAV programs to deliver ordinance to remote targets. The Navy
“Aerial Torpedo” and the Army “Kettering Bug” were tested with mixed results, but were never
used in combat (Keane, 2013). By 1924, the Navy Research Lab demonstrated the successful
flight of a radio controlled F-5L aircraft. The significance of this test flight was that the F-5L aircraft
was remotely controlled during all phases of flight (Keane, 2013). The Navy later sponsored
development of carrier based UAVs under the Joint Unmanned Combat Aircraft System program
(J-UCAS). With of over a decade of development, the J-UCAS program produced the TDR-1
assault drone (Gundlach, 2014). Considered the Navy’s first operational UAV, the TDR-1 flew
combat missions during World War II. With a 14.6 m wingspan and weighing 2676 kg, the TDR-1
could carry up to 907 kg of ordinance to a target 684 km away. With a cruising speed of 225
km/h, the flight endurance works out to be 3.04 hours. The TDR-1 was powered by twin Lycoming
O-435 internal combustion engines (ICE) each weighing 196.4 kg and producing 220 horsepower
or 164.1 kW (Interstate TDR, 2016). On board avionics systems included a television transmitter,
radar altimeter, and radio control receiver (Lee, 2013). A significant factor to the success of the
TDR-1 can be credited to the improved navigation and radio control electronics.
Today, there are a variety of fixed wing UAVs in both military and civilian service.
Introduced in 1996, the General Atomics RQ-1 Predator UAV provided aerial reconnaissance
capabilities to the U.S. military. In 2002, the addition of missile carrying capacity changed the
designation from RQ-1 to MQ-1. Military designation convention assigns the letter R for
reconnaissance, the letter M for multi-role, and the letter Q for unmanned. Further enhancements
to the Predator were released in 2005 changing the designation to MQ-1B which is still in service
today (“MQ-1B Predator”, 2015). With a wingspan of 16.8 m and a weight of 512 kg, the MQ-1B
has a range of 1427 km and a maximum flight endurance of 40 hours (“MQ-1 Predator/MQ-9
Reaper”, 2016). Powered by an ICE, the Rotax 914F four cylinder engine develops 115
UAV, a civilian UAV, is the Altair. As a joint development effort between NASA and General
Atomics, it is a modified version of a Predator B. Designed for high altitude scientific and
2
commercial research, it can reach a maximum altitude of 15.8 km. With a weight of 3175 kg and a
wingspan of 26.2 m, the Altair has a range of 12,441 km and a flight endurance of 32 hours
(Gibbs, 2015). Powered by the Honeywell TPE 331-10T turboprop engine, the power plant
develops 940 horsepower (700.9 kW) with a weight of only 174.6 kg (“TPE331-10 Turboprop
Engine”, 2006). The third example of a modern UAV is perhaps one of the most famous, the
Northrop Grumman RQ-4B Global Hawk. As the military designator indicates, it is an unmanned
reconnaissance system. Able to reach altitudes of 19.8 km, it is perhaps the highest flying UAV to
date. The Global Hawk weighs 6781 kg and has a wingspan of 39.9 m (“Northrop Grumman
Facts”, 2008). The Rolls Royce AE3007H turbofan engine produces 9.5 million newtons of thrust
with a power plant weight of 746 kg (“AE3007”, n.d.). The flight endurance is reported to be 35
hours with a range of 22780 km (“Northrop Grumman Facts”, 2008). The on board sensor system
combines electro-optical sensing, infra-red sensing, and synthetic aperture radar (“Northrop
Grumman RQ-4 Global Hawk”, 2016). Global Hawk is a paragon of how far UAV technology has
The current day market for UAVs is expanding, but the sector of small Unmanned Aerial
Systems (sUAS) has the largest growth potential. The designation of UAS instead of UAV
includes not only the airplane platform, but also the control and analysis equipment needed to
perform the mission objectives. The classification of “small” UAS is generally accepted as UAV
weight under 25 kg. The UAV is not valued on the novelty of a remote controlled aircraft, but
rather on the value of the mission it can perform. In other words, the value of the UAS is in the
capabilities it provides to the customer. For example, military UAVs perform intelligence,
surveillance, and reconnaissance (ISR) missions. The airplane platform includes imaging and
navigation equipment to record imaging and location data. The full UAS system includes the UAV
plus radio control equipment and analysis computers to interpret the data. The value of the UAS
is in the ability to gather imaging data with associated locations of a remote area. Similarly, the
value of a civilian UAS may be to provide thermal imaging data of crops over a group of large
3
fields for better farming. Within the aerospace industry, the small UAS sector is small, valued at
1.25 billion U.S. dollars in 2003 (NASA, 2006). However, the small UAS sector has also been
called the most dynamic of the aerospace industry due to its high growth potential. In 2015, ABI
research predicted the sUAS sector alone will exceed 8.4 billion dollars by 2018 (Manufacturing
Close-Up, 2015). Enabling this explosive growth, are advancements in electronics. Smaller,
lighter, and more capable systems can now be manufactured at a price low enough to interest
private sector customers. Previously, sUAS costs have been too high for many civilian
applications, but there is a trend of declining costs. Although sUAS price points are still above
conventional costs, sUAS solutions are capturing the attention of potential users. The expectation
is that sUAS solutions will soon be favorable for a wide variety of civilian applications.
As the cost of sUAS decrease, the variety of applications broadens and the user base
expands. During the early years of UAV development, demonstration of the technology was the
goal. The users were researchers and financiers. Although research and development costs were
high, they were acceptable to demonstrate proof of concept. After reliable service had been
proven, military and government sponsored programs were the only users; the only ones who
could afford the costs. For these end users, UAS applications were mainly military, namely ISR
and weaponry. As the technology matured, scientific programs were funded, such as the Altair
program. Continued development and innovation have now decreased the costs to the point
where commercial applications are possible. Scientific and civilian flight missions that are
considered dull, dirty, or dangerous are good candidates for sUAS applications. Areas of interest
such as earth science, land management, communications, and homeland security are being
considered (NASA, 2016). To continue the trend of increased UAS mission capability at a lower
cost, further improvements to the technology are being researched. Topics such as cost, flight
endurance, flight autonomy, sensor systems, and application data analysis are areas which will
4
1.4 UAV Hybrid Power Systems
Within the sUAS category, UAVs weighing 10 kg or less, often use electric motors as the
power plant instead of ICEs. There are many advantages to using an electric motor for small
scale UAVs. Electric motors have higher efficiency than their ICE counterparts. Electric motors
have efficiencies in the range of 80% to 90% while small ICE have efficiencies in the 10% to 20%
range. Since there are fewer moving parts, electric motors are more reliable and easier to
maintain. Electric motors have no emissions; zero carbon footprint. Also, electric motors have low
noise, vibration and thermal signature. The disadvantage of electric motors is electric power
storage. Due to limitations in battery technology, flight endurance is limited. With battery power
only, typical flight times are three hours or less. There is substantial interest in the UAV
Two approaches have been taken to improve flight endurance for electric powered, small
scale UAVs. The first approach is to mount photovoltaic cells on the aircraft to generate electric
power during flight. The second approach is to use fuel cells to convert stored hydrogen into
electric power. In this paper, the focus will be on the fuel cell approach. For many cases, the
power delivered by the photovoltaic cells or the fuel cell is not enough for the aircraft to takeoff or
climb. To compensate, lithium polymer batteries are combined with the alternative power source
(APS) to create a hybrid power system. In other words, battery power is used to augment power
delivery from the photovoltaic cells or fuel cell during the takeoff or climb segments of flight. An
example of a small scale, electric powered UAV using a hybrid power system is the FAUCON H2
from EnergyOr Technologies. With a wingspan of 3 m and a weight of 9 kg, the FAUCON H2
completed a flight of 10 hours and 4 minutes. Using a hybrid fuel cell and battery power system,
the flight lasted four times longer than using batteries alone (“EnergyOr fuel cell powered UAV
at the University of Johannesburg in 2009. A small Piper Cub model airplane was powered by a
custom 100 W fuel cell. The Piper Cub had a wingspan of 2.31 m and a weight of 5.35 kg. The
main objective of the project was to demonstrate that a small 100 W fuel cell could sufficiently
power the aircraft during level flight. To have enough power during takeoff and climb, lithium
5
polymer batteries were used. The radio control transmitter had one channel dedicated to
controlling a relay inside the aircraft to switch between batteries or the fuel cell. During takeoff
and climb, the batteries powered the electric motor. When the level flight or cruising portion of the
flight began, the power source was switched from batteries to the fuel cell. Although the flight
lasted only 9 minutes, the project goal was successfully demonstrated. That is, a 100 W fuel cell
can deliver sufficient power to maintain a small scale UAV in level flight (Furrutter, 2009).
Using fuel cells in combination with batteries in UAVs has a synergistic or complimentary
quality. Hydrogen fuel cells have a relatively high energy density but low power density. On the
other hand, batteries have a relatively low energy density but have a high power density. The fuel
cell can enhance flight endurance while the battery delivers high power when needed. Fuel cells
extend the flight endurance due to the high energy density of hydrogen. Compared to gasoline,
compressed hydrogen gas has more than 2.5 times the specific energy. In other words, for the
same amount of fuel weight, compressed hydrogen stores more than 2.5 times the energy than
gasoline. Additionally, fuel cells have an operational efficiency of roughly 3 times that of ICEs.
Ideally, the superior energy density of compressed hydrogen and the higher efficiency of fuel cells
can result in a 7.5 times or more increase in flight endurance. It is the potential increase in flight
endurance that has attracted interest and research in hybrid fuel cell and battery power systems
for small scale UAV applications. Actual flight endurance improvement is lower than the ideal
value due to practical engineering issues such as weight, power management, fuel cell system
Table 1
Even though fuel cells have the ability to extend flight endurance, from a practical
engineering point of view, there are still advantages and disadvantages to their use. Starting with
the advantages, fuel cells operate at a higher efficiency than ICEs. Typical proton exchange
6
membrane fuel cells (PEMFC) operate from 40% to 60% efficiencies while small ICEs have
efficiencies from 10% to 20%. With fuel cells, there are no moving parts, no vibration, and no
combustion; just direct energy conversion. Operation is quiet and dependable. There are no
carbon emissions, just water and heat as byproducts. The disadvantages include higher cost,
sensitivity to fuel contaminants, low power density, and a limited operational lifetime. Also, the
fuel cell requires supporting equipment for operation called balance of plant (BOP). The BOP
equipment consists of a fuel cell controller, a thermal management system, a water management
system, and possibly an air intake compressor; all of which consume power. Part of the fuel cell
output power may be used to run the BOP equipment, but a battery is needed to power the BOP
Power management of hybrid fuel cell and battery power systems in UAV applications
controls the power flow to and from both power sources. The goal of the power management
system (PMS) is to maximize flight endurance while keeping both fuel cell and battery within their
safe operational limits. Research in this area is ongoing and many different approaches have
been reported in literature. In general, the different power management approaches may be
classified into two broad categories, active and passive. Passive power management (PPM) are
cases where the fuel cell and battery are selected to be roughly the same voltage and are
connected directly in parallel. The combined power is then sent to the electric motor to power the
aircraft. The advantage of PPM is that there are no extra electronics to carry along in the UAV;
weight is kept to a minimum. The drawback of PPM, is that there is no control of power flow in the
system. There is no mechanism to ensure the fuel cell and battery operate within their limits. If
there is voltage mismatch between the fuel cell and battery, then energy may be wasted by
unmanaged power flow between the two sources. In these cases, the two power sources are said
to be in contention. The second category, active power management (APM), uses electronics to
control power flow in the system. The fuel cell and battery are kept within operational limits. The
downside to APM is that there are supervisory electronics to carry in the UAV, thus increasing the
weight and consuming power. An APM system may be as simple as using a relay to control which
7
source delivers power to the motor. The Piper Cub example cited earlier is a good example of a
simple APM approach. The beneficial feature of the Piper Cub APM approach was that the weight
and power consumption of the APM was minimal. Only a relay and one channel from the radio
control were used. The disadvantage was that the remote operator had to know when to switch
between power sources. An APM system may be very complicated such as using multiple DC-to-
DC converters under microprocessor control. In these cases, control of power flow can be closely
managed and optimal results attained. The cost for such control is extra weight, cost, and power
consumption.
The PPM electrical architecture centers on an electrical bus which serves to interconnect
all power sources and power loads. The bus is simply two wires; one positive and one negative.
The negative bus wire is considered ground for the power system; the electrical reference
potential to which all other power system voltages are referenced to. The positive bus wire is at
the power system voltage. For example, a 12 V power system means that the positive bus wire is
at plus 12 V relative to the negative bus wire. The point is that a voltage is always the electrical
potential difference between two points; in this case the bus positive wire relative to the bus
negative wire. The PEMFC positive terminal, the battery positive terminal, and the electronic
speed controller (ESC) positive input terminal are connected to the positive bus wire. Similarly,
the negative terminals of the three components are connected to the negative bus wire. In effect,
the two power sources and the power load are connected in parallel which constrains the voltage
across them to be equal. The power load is the electronic speed controller. The ESC is an
electronic device that drives the brushless DC motor (BDCM) to a constant speed. It receives its
input power from the bus and speed commands from a separate electronic input connected to the
radio control receiver. When set to zero, the ESC draws no power from the bus and sends no
power to the BDCM. In reality, the ESC still draws a small amount of power from the bus at a zero
setpoint; enough power to run its internal electronics (quiescent power). At this point, there is zero
electrical load on the power bus (neglecting the load quiescent power). Ideally, the PEMFC and
battery currents should be zero; unfortunately, they are not. A small mismatch in the PEMFC
voltage and the battery voltage results in a large current between them. The current is a result of
8
the parallel connection between the power sources which constrains the terminal voltages to be
equal. The end result is that the current between the power sources will increase until the power
source voltages are the same. Depending on the magnitude of the power source voltage
mismatch, the current flow between them can be large which results in a significant waste of
energy. The two power sources are in contention. When designing a PPM system, the voltage
mismatch between the PEMFC and battery pack is minimized in order to minimize wasted
energy. As the ESC setpoint value increases, electrical power is drawn from the bus to the ESC
which in turn sends electrical power to the BDCM. The load power drawn from the bus continues
to increase until the BDCM is rotating at the ESC setpoint. The power sources now deliver power
to the ESC and to each other. As the load on each power source increases, the operating point
on the voltage versus current curve (VI-curve) for each power source changes. A power system
balance is reached where the load receives its required power and the two sources split the
power load according to the equilibrium point reached on their respective VI-curves. At higher
load currents, the power sources have an opportunity to equalize their terminal voltages by
establishing a ratio of current delivered to the load. If the source VI-curves are somewhat close,
the power exchange between them can become small since their terminal voltages can equalize
by splitting the load current. In practice, the PPM scheme is used and the energy wasted at low
load currents is tolerated. The advantage of the PPM scheme is that no supervisory power
management electronics are used, thus minimizing complexity, cost, and weight. The
disadvantages of the PPM scheme is that energy is wasted by power exchange between the
sources and there is no control over the battery state of charge (SOC) and the fuel cell operating
point.
The APM electrical architecture is different from the PPM case in that it controls power
flow from the sources. The goal is to minimize wasted energy at the power sources and maximize
the energy delivered to the load. There are many ways to implement an APM system. The Piper
Cub example is a simple case, where a relay is used to connect either the battery or the fuel cell
to the bus, but never both at the same time. During takeoff and initial climb of the flight profile,
only the battery is connected to the bus, delivering high power to the load. When the cruise phase
9
of the flight profile is reached, the relay switches over to the fuel cell. The battery is disconnected
from the bus and the fuel cell is connected to the bus. The fuel cell then delivers a reduced power
level to the load, but no power flows from the fuel cell to the battery. In this manner, the relay
electrically separates the two power sources and there is never contention between the sources.
one or more DC-to-DC converters as a key component in APM electrical architecture. For a
power source connected to the electrical bus through a DC-to-DC converter, the amount of power
delivered to the bus can be controlled by the output voltage of the converter. For example,
consider a hybrid fuel cell and battery power system. The fuel cell is connected to the bus through
a DC-to-DC converter while the battery is connected directly to the bus. The APM controller can
adjust the DC-to-DC converter output voltage which effectively varies the fuel cell voltage
presented to the bus. If the controller keeps track of the battery SOC, it can then use this
information to maintain the battery SOC within safe limits. At the same time, adjusting the DC-to-
DC converter output voltage effectively limits the load current of the fuel cell. A disadvantage to
the use of DC-to-DC converters between a power source and the bus is the weight of the
converter. High power DC-to-DC converters are heavy, and significantly increase the weight of an
APM system. In UAV applications, excessive weight due to the APM system detracts from aircraft
performance.
storage device. Current research investigates the application of alternative energy sources such
as fuel cells or photovoltaic arrays to UAV power systems. In this study, photovoltaic power is not
pursued and the focus is on fuel cell and battery energy sources as applied to UAV power
systems. Batteries alone can supply enough electric power for UAV flight, but have a limited flight
time; typically under 3 hours. Fuel cells present an alternative to batteries and have the potential
for longer flight times. Unfortunately, to handle all electric power requirements for UAV flight, a
fuel cell must be sized so that its power rating can handle the maximum power requirement. For
UAV flight, power consumption can be three to five times higher during takeoff than during regular
10
cruising conditions, but the higher power is only required for a short period of time; roughly 2% to
4% of the total flight time. The result is that the fuel cell size is much larger and heavier than
necessary for 96% to 98% of the flight time. The combination of fuel cell and battery for UAV
power systems allows the fuel cell to be properly sized for the majority of flight time while the
battery can assist with power delivery during takeoff. A hybrid power system configuration
consisting of a fuel cell combined with a battery can significantly improve flight endurance and still
meet all power requirements for UAV flight. The first research goal of this study is to create
simulation models of the battery, fuel cell, and electrical load. The simulation environment
employed for this work is Matlab and the simulation components are written in the Simscape
language. Simscape runs in the Simulink simulator which is part of Matlab. Simscape is a new
addition to the Simulink toolbox and is used to create simulation models of physical systems.
Active power management of a hybrid power system controls the electrical load sharing
between the fuel cell and battery. The power required by the load must be met at all times during
the flight profile and the fuel cell and battery must be maintained within their safe operating limits.
In many active power management approaches, a DC-to-DC converter is used in the main power
path to the load. DC-to-DC converters with large power ratings are heavy and thus detrimental to
overall UAV performance. An active power management system that does not use a DC-to-DC
converter in the main power path will lower power system weight and thus improve overall UAV
performance. Typical of UAV hybrid power sources, the battery is of the Lithium Ion (Li-Ion) type,
or more specifically, Lithium Polymer (LiPo). For such batteries, the SOC must be kept within
lower and upper limits or damage to the battery may result. In cases where the SOC remains
outside the safe operating zone, the battery may spontaneously ignite and cause a hazardous fire
condition. A popular type of fuel cell used in UAV hybrid power systems is the PEMFC. The
current drawn from the PEMFC must be controlled so that the maximum current rating is never
exceeded. If current drawn from the PEMFC exceeds the maximum rating, then the fuel cell will
shut down as a safeguard against permanent damage. The second research goal of this study is
to develop an active power management scheme that does not use a DC-to-DC converter in the
main power path of the hybrid power system and maintains both battery and fuel cell within their
11
safe operating zones. A simulation model of this hybrid power management controller is then
created in Simscape.
To create a complete model of the UAV power system, the components are combined into a
working simulation model. The battery, fuel cell, electrical load, and power management controller
models are combined in a top level Simscape simulation model to evaluate the flight performance
of the power system. For a given flight plan, a power usage profile or flight profile can be
generated. To determine the amount of hydrogen fuel required for the given flight profile, the
simulation is run. Simulation results can then assess power system performance by inspecting
battery SOC, fuel cell current, and the power split between fuel cell and battery during flight. A
feature of the hybrid power management controller is the ability to change the fuel cell current
during flight. When this feature is used during cruise conditions, the battery may contribute part of
the load current thus lowering the amount of current the fuel cell supplies. Lower current from the
fuel cell extends the hydrogen fuel supply and therefore the flight time as well. This feature of the
controller is named current assist and can be used to extend flight endurance of the UAV. The
third research goal of this study is to determine the amount of hydrogen fuel necessary to
complete a six hour flight profile. The hydrogen consumption is to be determined with current
assist turned off and also with current assist active. Reviewing battery SOC and fuel cell current
during flight will establish if the controller has successfully maintained the battery and fuel cell
During short flights, a UAV may have excess hydrogen available. For these scenarios,
recharging the battery during flight may be advantageous. For example, if a UAV is trying to climb
to maximum altitude, the battery may be a limiting factor. Using excess hydrogen to recharge the
battery inflight may allow higher altitudes to be reached. Another benefit of inflight battery
recharge is that during ground maintenance, a recharged battery may shorten service time or
allow for shorter turnaround times for back to back missions. A battery charger model is
developed and integrated into the UAV power system model. The fourth research goal of this
study is to demonstrate inflight battery recharge by simulation of the given flight profile with the
12
battery charger active. The battery SOC should have an upper limit and the battery charge should
For UAV hybrid power system design, selection of the fuel cell, battery, and hydrogen storage
tank will determine the UAV flight endurance and power system weight. Although the total weight
of the power system can be determined during the design phase, the total flight time may be
difficult to determine. For a given flight profile, simulation results can be used to derive a simple
formula that applies to the particular UAV case. The formula can then be used to evaluate power
system weight versus potential flight time to help choose the best combination of components.
The fifth and final research goal of this study is to derive such a formula for the ASU Composite
UAV using a given six hour flight profile. The resulting formula is then used to create a plot of
power system weight versus flight time for combinations of commercially available fuel cells,
13
CHAPTER 2
LITERATURE SURVEY
For electric powered UAVs, using hydrogen fuel cells to extend the total flight time has
increased interest in the development of experimental UAVs using fuel cell power systems has
occurred. For example, the UAV manufacturer AeroVironment Inc. developed a micro air vehicle
named the Wasp. With a wingspan of only 9 inches and a weight of 7 ounces, it carried a global
positioning system (GPS) module and two cameras for aerial reconnaissance missions. The
Wasp was powered by a lithium ion battery and set an endurance record of 100 minutes during
its initial flight in 2002 (Fuentes, 2005). By 2003, the first fuel cell powered UAV was developed
by AeroVironment. Based on the earlier Wasp design, the Hornet replaced the lithium ion battery
with hydrogen fuel cells. The fuel cells had the potential of doubling the flight endurance (Moffitt,
2010). A unique design feature of the Hornet was that the fuel cells were integrated directly into
the structure of the wings. The fuel cells had two purposes, to serve as part of the wing structure
and to supply power to the electric motor and onboard electronics. The intended benefit of
exposing the fuel cells at the top wing surface was to have plenty of airflow to supply oxygen to
the chemical reactions inside the fuel cells (Jefferson, 2002). Unfortunately, the airflow over the
fuel cells proved to be excessive during the Hornet’s maiden flight in 2003. The flight time was
reduced to only five minutes due to failure of the fuel cells. The problem was that a specific range
of water content inside the fuel cell membranes needed to be maintained, but the excessive
airflow caused the fuel cell membranes to become dehydrated (Moffitt, 2010).
14
Figure 1. Hornet UAV (www.avinc.com/images/uploads/general/7/hornet2-1_bg.jpg)
The U.S. Naval Research Laboratory (NRL) has an ongoing program to develop fuel cell
powered UAVs. Founded in 1923 by T. Edison, the Navy’s corporate research laboratory started
out with only two divisions, radio and sound (“NRL History”, 2016). Early accomplishments of the
NRL include the successful flight of the radio controlled aircraft F-5L (Keane, 2013). Since the
early years, NRL has grown to encompass 18 general areas of research with 21 divisions (“NRL
Fact Book”, 2014). The alternative energy section (6113) of the NRL chemistry division is one
group researching fuel cell powered UAVs (“DOE FCTO”, 2016). In November of 2005, NRL
completed the first test flight of a small, fuel cell powered UAV named the Spider Lion. With a 95
W hydrogen fuel cell, the 1.71 kg aircraft flew for 3 hours and 19 minutes from 15 grams of
The hydrogen fuel cell used in the Spider Lion was manufactured by Protonex Technology Corp.
of Southborough, Massachusetts (“USAF contract for Protonex UAV power”, 2006). The next
NRL program to advance fuel cell technology in UAV applications was the Ion Tiger program. The
goal of the Ion Tiger program was to design and test a UAV capable of carrying a 2.26 kg payload
for a flight time of 24 hours. A custom airframe was necessary since no off the shelf airframes
15
had enough fuselage volume to accommodate the fuel cell and compressed hydrogen storage
tank while still having a high enough lift to drag ratio. Instead, the airframe was designed at NRL
using carbon fiber and Kevlar construction. The fuel cell, built by Protonex, was a 550 W, PEMFC
weighing only 1 kg. The hydrogen storage method used compressed hydrogen gas stored in a
carbon overwrapped aluminum pressure vessel (Swider-Lyons, 2011). Finished in the fall of
2009, the end result was an aircraft weighing 13.8 kg with a wingspan of 5.17 meters, capable of
carrying a 2.26 kg payload for a 24 hour flight. The first test flight of the Ion Tiger occurred in
October 2009 under windy conditions. The Ion Tiger flew for 23 hours and 19 minutes carrying a
1.81 kg payload. The total flight time fell slightly short of the 24 hour goal since the hydrogen burn
rate was higher than planned due to the windy conditions. In November 2009, under calmer wind
conditions, the Ion Tiger flew for 26 hours and 1 minute carrying a 2.26 kg payload (Swider-
Lyons, 2011). Post flight analysis revealed that the October flight under windy conditions had an
average propulsion power of 326 W while the November flight (mild conditions) had an average
propulsion power of 314 W. In both cases, 500 grams of hydrogen were consumed, but the total
energy delivered during the test flights were 7600 Wh for the October flight and 8160 Wh for the
In May 2013, the Ion Tiger reached a longer flight endurance of 48 hours and 1 minute using
liquid hydrogen storage instead of compressed gas storage (“Ion Tiger Fuel Cell Powered UAV”,
2016).
16
2.2 Hybrid Powered UAV Examples
A power system is said to be hybrid when two or more power sources are combined to
deliver power to the same load. For example, a lithium ion battery combined with a fuel cell
configured such that both sources deliver power to the same electric motor is considered a hybrid
power system. For UAV applications, a hybrid power system design can significantly reduce the
size of the fuel cell. For UAV cases where the aircraft takes off from the ground under its own
power (runway), the power required during the takeoff phase of flight is much higher than the
power required during airborne cruising conditions. A rough estimate of the ratio of required
takeoff power to minimum required cruising power is about four to one. The four to one ratio
depends upon the design of the aircraft and is a ratio estimate for a fixed wing UAV developed at
Arizona State University (ASU) (C. Nam, personal communication, September 17, 2015).
Although this ratio may vary for each UAV case, it is a helpful estimate to evaluate hybrid
powered UAV design. For example, a UAV that requires 100 W of power during cruising
conditions (steady state) would need 400 W to take off from a runway. A hybrid power system
can be employed in such cases by having the battery augment the fuel cell power during takeoff.
The fuel cell can then be used to power the UAV during the cruising phase of flight without help
from the battery. In effect, the hybrid approach reduces the required fuel cell size from 400 W to
100 W. During the takeoff phase of flight, the fuel cell delivers 100 W of power combined with 300
W of power from the battery. During the cruising phase of flight, the fuel cell continues to deliver
100 W of power and the battery delivers no power at all. In this manner, the synergistic
combination of fuel cell and battery maximizes flight time while minimizing the size and weight of
the power system. An example of a hybrid fuel cell and lithium ion powered UAV was developed
in 2010 by the National Cheng Kung University in Taiwan. Named the Grey-faced Buzzard, the
30 kg fixed wing aircraft was powered by a 1 kW fuel cell and 5.4 Ah lithium ion battery.
Combined, the fuel cell and battery deliver 2.5 kW of electric power to the motor at takeoff. With a
wingspan of 3.4 m and an electric motor rated at 4 kW, there is enough lift and propulsion to carry
an 8 kg payload. The test flight of the Grey-faced Buzzard lasted 15 minutes covering a distance
of 30 km. Although the test flight of the Grey-faced Buzzard was short, it was a successful
17
demonstration of the hybrid power system (“Fuel cell/battery hybrid UAV takes off in Taiwan”,
2010). In this case, the ratio of total power to fuel cell power is 2.5 to one. If the ratio for takeoff
power to cruising power is four to one, then a maximum power of 2.5 kW at takeoff yields a
required minimum cruising power of 625 W. The actual ratio may differ, but assuming four to one
is within 25% of the actual ratio, the Grey-faced Buzzard has an excess fuel cell power capacity
of 160 W or more during cruising flight. The excess fuel cell power may be used to power the
onboard electronics, power the fuel cell BOP equipment, or to increase propulsion power during
Two commercially built, hybrid powered UAVs are the FAUCON H2 from EnergyOr
Technologies and the Stalker XE from Lockheed Martin. Both UAVs use fuel cell and lithium ion
battery hybrid power systems, but the type of fuel cell is different in each. The FAUCON H2
utilizes a proton exchange membrane fuel cell (PEMFC) while the Stalker XE has a solid oxide
fuel cell (SOFC). The PEMFC is a low temperature fuel cell; operating at a relatively low
temperature range, from 30oC to 100oC. On the other hand, a SOFC is considered a high
temperature fuel cell; operating in a temperature range from 500oC to 1000oC (Larminie & Dicks,
2003). The PEMFC uses compressed hydrogen gas for fuel while the SOFC in the Stalker XE
consumes propane (LPG) fuel. Both aircraft employ lithium polymer batteries to augment the fuel
cell power when needed. These two UAVs are an interesting comparison since they are of similar
size but use different fuel cell technologies. In addition, neither UAV is considered an
experimental aircraft; both are operational, ruggedized sUAS; robust enough for commercial
18
missions. Both systems have a full complement of onboard avionics along with a ground control
station. The FAUCON H2 has a wingspan of 3 m, a weight of 9 kg, and a payload capacity of 1 kg
(“EnergyOr fuel cell powered UAV reaches 10 h flight endurance”, 2011). The Stalker XE has a
wingspan of 3.66 m, a weight of 7.5 kg, and a payload capacity of 2.5 kg (“Stalker UAS”, 2016). In
2011, the FAUCON H2 completed a demonstration flight lasting 10 hours 4 minutes, then
performed a successful autonomous landing (“EnergyOr fuel cell powered UAV reaches 10 h
flight endurance”, 2011). Also in 2011, Lockheed Martin reports the Stalker XE to have an 8 hour
flight endurance after numerous flight tests (“Lockheed Martin ruggedized UAS uses AMI fuel cell
power”, 2011). Both UAV manufacturers report the hybrid power system increases flight
Two UAVs constructed at the university level are the Piper Cub UAV from the University
of Johannesburg and the Composite UAV from ASU. Constructed in 2009, the Piper Cub was
built to demonstrate that a 100 W fuel cell provides enough power for steady state flight. With a
wingspan of 2.31 m and a weight of 5.3 kg, the Piper Cub flew at cruising conditions for 9
minutes. During takeoff, the Piper Cub was powered by a lithium polymer battery (LiPo) and when
it reached steady, level flight (cruising conditions), power was manually switched over to the fuel
cell by remote operation on the radio controller (Furrutter, 2009). Internally, one of the radio
control channels determined the state of an electrical relay which connected either the battery or
the fuel cell to the electric motor. At no time was the fuel cell and battery both delivering power to
the electric motor simultaneously. The PEMFC used was the Horizon Fuel Cell Technologies H-
100 and the LiPo battery type is not reported (Furrutter, 2009). In 2012, ASU constructed a
Composite UAV based on the AIAA 2012 “Design Build Fly” competition specifications. Designed
19
and built by undergraduate students, the airframe was a custom design constructed out of carbon
fiber composite materials. The Composite UAV has a 2.03 m wingspan, a weight of 5.5 kg, and a
payload capacity of 2 kg (Krauch et al., 2012). The aircraft was powered by three electric motors;
one mounted on the front nose of the fuselage and two mounted on the trailing edges of the main
wing. The initial test flight in December 2012 was under LiPo battery power only. Unfortunately,
the flight was cut short by crashing soon after takeoff. Later, a second Composite UAV was
constructed and has become the centerpiece for the continued effort at ASU Polytechnic campus
to develop a hybrid powered UAV. In the fall semester of the following year, a hybrid power plant
study was completed at ASU Polytechnic campus (Monaco, 2013). The goal of this study was to
increase the flight time of the Composite UAV from an estimated 25 minutes using LiPo batteries
only, to 6 hours using a hybrid fuel cell and LiPo battery power system (Monaco, 2013). The work
presented in this thesis is a natural continuation to further develop the hybrid power system
architecture proposed by Monaco. To compare the Composite UAV and Piper Cub UAV, both
airframes are of roughly the same size and weight. Both aircraft are of similar design; that is both
airframes have a conventional high wing placement of the main wing and a conventional tail
design. Due to this similarity, it is reasonable to expect that the 100 W fuel cell size used in the
Piper Cub UAV will be somewhat close to the minimum fuel cell size required in the Composite
Figure 6. Piper Cub UAV (left) (www.google.com), ASU Composite UAV (right)
(Nam, C., lecture notes, 2015)
20
Figure 7. ASU Composite UAV CAD Assembly (left) (Nam, C., lecture notes 2015),
ASU Composite UAV Disassembled (right) (Nam, C., lecture notes 2015)
Figure 8. ASU Composite UAV Test Flight, December 2012 (Nam, C., lecture notes 2015)
Among the different portable, secondary battery types available, lithium ion battery
technology offers the highest energy density. Compared to other rechargeable battery types such
as nickel-cadmium (Ni-Cd), nickel metal-hydride (Ni-MH), or lead acid batteries, lithium ion
batteries are the best choice for UAV power systems. Factors such as energy density, cell
voltage, weight, volume, and discharge rate are primary considerations for battery type selection.
Secondary factors such as cost and charge retention are attributes to consider as well but do not
have disparities large enough to disqualify lithium ion technology for typical sUAS applications.
21
Lithium is the preferred anode material due to its high standard reduction potential (-3.05 V) and
its low weight (0.534 g/cm3) resulting in higher energy densities of around 150 Wh/kg
(Madakannan, lecture notes 2014). With a nominal cell voltage of 3.7 V, lithium ion batteries
obtain higher battery voltages with fewer cells connected in series. Higher voltage batteries need
to supply less current to deliver the same amount of power. When a battery delivers less current,
then the amount of stored active chemicals within the battery lasts longer. The result is a relaxed
constraint on the amount of active chemical material stored within the battery to deliver a
specified amount of energy. Due to the higher cell voltage of lithium ion batteries, fewer cells are
required to obtain a desired battery voltage. Less active chemicals need to be stored to reach a
specified energy storage goal. The end result is lower battery weight and volume for lithium ion
batteries as compared to other options. The benefit of lithium ion batteries is clarified when
considering the weight and size required to deliver 100 Wh of energy to a load (e.g. an electric
motor). Discharge rate of early lithium ion batteries used to be a restricting factor for high
discharge rate applications, but the technology has improved over time to where this is no longer
the case. Zinc-Air batteries have an even higher energy density than lithium ion batteries, but
they cannot be used in UAV applications since they have very low discharge rates. Cost of lithium
ion batteries can be three to five times that of competing technologies. Although battery cost is a
factor, lithium ion battery cost is still low enough to be the popular choice of battery type in both
Table 2
Battery Type Comparison (Madakannan, lecture notes 2014, Linden & Reddy, 2002)
Battery Specific Energy Volumetric Energy Nominal Cell 100 Wh 100 Wh
Type Density (Wh/kg) Density (Wh/L) Voltage (V) Weight (kg) Volume (L)
Lead Acid 35 70 2.00 2.86 1.43
Ni-Cd 35 100 1.20 2.86 1.00
Ni-MH 75 240 1.35 1.33 0.42
Li-Ion 150 400 3.70 0.67 0.25
Zinc-Air 220 442 1.65 0.45 0.23
22
Figure 9. Battery Type Volumetric Energy Density versus Specific Energy Density
(Madakannan, lecture notes, 2014)
Although the chemistry and construction of lithium ion batteries have many varieties, the
basic operating principles are the same. During discharge, lithium ions travel from the anode to
the cathode passing through the electrolyte layer. The process is completely reversible so that
the battery can be recharged. During charge, lithium ions travel from the cathode back to the
anode. Typical anode electrode material consists of lithiated carbon, LixC6. The layered carbon
anode material acts as a host matrix (crystal lattice) for lithium ions. In other words, lithium ions
can be freely inserted into and extracted from gaps between layers of the carbon material crystal
lattice; a process known as intercalciation. During discharge, lithium ions are extracted from the
anode crystal lattice. When a lithium ion is extracted, the anode material is oxidized resulting in a
free electron which is available to be delivered to the external electrical load. The lithium ion then
travels through the electrolyte to the cathode where it is inserted into the cathode crystal lattice.
When the lithium ion is inserted into the cathode crystal lattice, the cathode material is reduced
thus absorbing an electron from the external electrical load. Typical cathode electrode materials
vary; a common one being LiCoO2. Since lithium ions may be inserted into and extracted out of
the anode and cathode electrode materials, these materials are named lithium insertion
compounds. This process of intercalciation is the core principle behind every type of lithium ion
battery. The electrolyte is the layer separating the anode and cathode. An electrolyte material
must be an insulator to electrons and a good conductor of lithium ions. Electrolytes may be a
23
liquid, a gel, or a solid, such as a polymer or a ceramic. In many commercial lithium ion batteries,
the electrolyte is a liquid, consisting of an organic solvent with inorganic solutes. A typical liquid
electrolyte may consist of a Tetrahydro furan (THF) solvent with a LiPF6 solute dissolved at a
A Lithium Polymer (LiPo) battery is a type of lithium ion battery, but the name is used in
two different ways. Within the scientific and industrial research communities, the name lithium
polymer battery means a lithium ion battery where the electrolyte is a polymer. As shown in figure
9, lithium polymer batteries have even higher energy densities than lithium ion batteries with
liquid electrolytes. These types of LiPo batteries are widely researched, but are not commonly
available commercially. The second meaning of lithium polymer battery is used by the battery
industry. In this context, a lithium polymer battery is a lithium ion battery with a liquid electrolyte,
but it is packaged in a prismatic, flexible package. The polymer in this case refers to the casing of
the battery and not the electrolyte. LiPo batteries that are commonly available to consumers are
of the second type; that is a lithium ion battery with liquid electrolyte packaged in a flexible,
For small model, electric aircraft applications, hobbyists typically use lithium polymer
batteries (LiPo). The same battery type is a good match for sUAS applications as well. The hobby
aircraft industry has well established that lithium polymer batteries are the best choice to minimize
weight while maximizing flight endurance. The rating system for these batteries has two key
parameters, the battery configuration and the maximum discharge rate. The battery configuration
consists of a four character designator xSyP. The x placeholder is a number typically from 1 to 5.
It is the prefix to the S designator which stands for the number of cells in series. For example, a
24
4S designator indicates that four lithium ion cells are in series within the battery pack. At 3.7 volts
per cell, a 4S battery pack will have a battery pack voltage of 14.8 volts. The y placeholder is a
prefix to the P designator, which indicates the number of cell strings in parallel. For example, a 2P
designator indicates that two lithium ion cell strings are in parallel. The more cell strings in
parallel, the greater the battery pack capacity. To clarify, a 4S2P designator indicates that there
are four lithium ion cells in series within a string, and that two strings are present in the battery
pack connected in parallel. Many times, the yP portion of the designator is left out, meaning that
there is only one cell string present in the battery pack. For example, a LiPo battery pack
designator of 4S indicates that there are 4 lithium ion cells connected in series and that there is
only one string present in the battery pack. The second key parameter for LiPo battery packs is
the maximum discharge rate. Measured in units of C (capacity), the discharge rate is listed as a
multiple of C. For example, if a battery pack has a capacity of one amp-hour, then a maximum
discharge rating of 20C indicates that the battery pack is capable of delivering an instantaneous
current of 20 amps.
There are a variety of methods to model lithium ion battery packs for purposes of system
simulation. Reported in literature are empirical, electrochemical, electric circuit, and abstract
model types (Tsang, 2010). The goals of the system level simulation will determine which battery
model is appropriate. At the most basic level, common to all models, is the prediction of battery
voltage as a function of current. An electrical circuit for such a battery model consists of an ideal
voltage source and a series resistor. The ideal voltage source represents the open circuit battery
voltage while the series resistor represents the internal resistance of the battery. In the equation 4
below, Vbat is the terminal voltage, Eo is the ideal voltage source value, R is the internal resistance
of the battery, and the independent variable i represents the battery current.
Vbat Eo i R (4)
With zero current draw, the battery voltage measured across the positive and negative terminals
of the battery (terminal voltage) is equal to the ideal voltage source value. As current drawn from
25
the battery increases, the terminal voltage drops linearly from the ideal voltage value. For this
simple battery model, only the internal resistance of the battery is accounted for; key parameters
such as total battery capacity and battery SOC are not included. The limitation of this model is
that the battery will run forever; power will continue to be delivered regardless of battery capacity,
the initial SOC, and the history of delivered power. A more realistic model will include the limited
electrical capacity of the battery. For example, a one amp-hour battery capacity states that a
battery, starting out at 100 percent SOC, is able to supply one amp of current to an electrical load
for one hour. During this discharge process, the battery voltage will decrease until it reaches a
final value close to zero. At this point, the battery is completely discharged and can no longer
deliver any electrical power. A more suitable model for the system level simulation purposes of
this study is presented by Olivier Tremblay et al. and is discussed further in section 3.2 (Tremblay
batteries, it is fundamentally different than a battery. Both a battery and a PEMFC use chemical
oxidation and reduction to produce electrical energy. The main difference between them is the
source of chemical reactants. In a battery, the chemical reactants are stored within the battery,
thus limiting the electrical capacity. In a PEMFC, the chemical reactants are external; fed to the
fuel cell during operation. The PEMFC ideally has an infinite capacity; as long as hydrogen fuel
and air are continually fed to the fuel cell, it will continue to generate electrical power. Of course a
practical PEMFC has endurance limitations due to degradation during operation, but the
mechanisms that degrade PEMFC operation over time are different than the limited amount of
chemical reactants that limit battery capacity. The physical construction of a PEMFC is quite
different from a battery. The PEMFC accepts input chemical reactants in gaseous form while
most batteries store chemical reactants as a solid or liquid. The PEMFC must also have a
mechanism to purge water produced from the chemical reaction while batteries do not have this
issue. A PEMFC also has to regulate humidity of the input gaseous reactants to maintain
membrane hydration during operation while batteries do not. PEMFC must also manage its
26
temperature during operation. If the PEMFC runs too hot, then the membrane will become
dehydrated and the fuel cell will stop functioning. Batteries in general do not have to regulate their
temperature during operation. There may be some esoteric cases where battery temperature may
have to be regulated, but these cases are not the norm. Typically, a PEMFC has a
fuel cell controller. All of the equipment to maintain proper conditions of the PEMFC during
operation are grouped together and labeled balance of plant (BOP) equipment. PEMFC BOP
equipment include the fuel cell controller, temperature management equipment, water content
management equipment, and possibly compressors to boost gas intake pressures. Clearly, a
the electrodes. Hydrogen gas (fuel) is fed to the anode. At the anode, gaseous, diatomic
hydrogen molecules are oxidized resulting in two hydrogen ions and two electrons.
The hydrogen ions pass through the electrolyte to the cathode. Same as battery electrolytes, the
PEMFC electrolyte must be an insulator to electrons while being a good ionic conductor. The free
electrons are available to the external circuit and their flow from the anode to the cathode must
pass through the external electrical load and not through the PEMFC electrolyte. At the cathode,
electrons return to the fuel cell from the external electrical load and combine with the hydrogen
ions and oxygen gas extracted from the air fed to the cathode.
1
Cathode reduction: 2e 2H O2 H 2O (6)
2
The end result is that hydrogen gas (the fuel) and air (the oxidizer) are combined to yield water,
1
Overall reaction: H2 O2 H2O (7)
2
27
At the core of the physical construction of a PEMFC is the membrane electrode assembly
(MEA). The central component of the MEA consists of a thin sheet of polymer material which acts
as the fuel cell electrolyte; also called a membrane. Typically, the membrane is a special polymer
developed by the DuPont Corporation named NafionTM. Similar to Teflon, Nafion has the special
property of conducting positively charged hydrogen ions while acting as an insulator to electrons.
Typical thickness of the Nafion membrane ranges from 18 um to 25 um. Working outwards from
the center, on both sides of the membrane are the catalyst layers; one side being the anode
catalyst layer and the other being the cathode catalyst layer. The catalyst layers are essentially
the same on both sides, consisting of fine platinum particles dispersed over a supporting structure
of carbon powder immersed in a polymer binding agent. The details of the catalyst layers are
critical to efficient fuel cell operation. The thickness of the catalyst layers range from 5 um to 30
um. Moving outwards from the catalyst layers are the gas diffusion layers (GDL); one on the
anode side and the other on the cathode side. The GDLs are constructed from porous carbon
paper or cloth which allow the reactant gasses to diffuse through them. Typical thickness of the
GDLs range from 174 um to 450 um. Next, the outer layers of the MEA are the electrode plates;
one anode plate and one cathode plate. The electrode plates are made of either a graphite
composite material or metal. The electrode plates provide physical support for the entire MEA
and are the outer plates used to compress the sandwich structure of the MEA. Among the other
functions implemented by the electrode plates, they provide a conduction path for electrons in
and out of the MEA and form the gas flow channels to deliver reactant gas to the GDLs
28
Figure 10. Membrane Electrode Assembly
A more detailed discussion of the MEA reveals many of the material requirements of
each layer which in turn illustrates the need for the multi-functional role that each layer provides.
At the center of the MEA is the Nafion membrane. Constructed from a sulphonated fluoro-
polymer material, the membrane requires a specific range of water content to conduct hydrogen
ions (Larminie & Dicks, 2003). Note that a positively charged hydrogen ion is a single proton, and
that the membrane is designed to be a proton conductor when properly hydrated. This is the
basis for the PEMFC name, proton exchange membrane fuel cell. One of the engineering
challenges of PEMFC design is to keep the membrane properly hydrated during operation. The
Nafion material has a fixed ratio of hydrophobic and hydrophilic properties such that it retains the
right amount of water if available. In other words, the Nafion membrane can dry out if not enough
water is present at its surfaces, but it will not become too hydrated if an excess amount of water
is available. To keep the membrane hydrated, the input hydrogen gas is humidified. Water vapor
carried by the hydrogen gas diffuses through the anode GDL and anode catalyst layer to the
membrane. On the cathode side, the cathodic reaction produces water at the cathode catalyst
layer. To prevent excess water from blocking the pours of the cathode catalyst layer and the
29
cathode GDL (flooding), it must be removed. Water diffuses outwards from the cathode catalyst
layer, through the cathode GDL, to the cathode plate, where it evaporates to the air flow. In
practice, the water management problem is more complicated, but this simplified explanation
keeps the description brief. Heat generated by the chemical reactions at the anode and cathode
catalyst layers must also be removed to keep the operating temperature of the fuel cell below
100oC. The reason for the upper temperature limit is to keep the membrane hydrated. An
operating temperature above 100oC will cause the membrane to dry out. Heat is conducted away
from the catalyst layers by the GDLs and ultimately dissipated by the electrode plates. Electrons
produced at the anode catalyst layer must be conducted through the anode GDL to the anode
plate to supply power to the external circuit. On the cathode side, electrons return to the cathode
plate from the external circuit and are conducted through the cathode GDL to the cathode catalyst
layer. Gas transport must occur on both the anode and cathode sides. Reactant gasses must
diffuse from the flow fields at the electrode plates, through the GDLs to reach the catalyst layers.
The catalyst layers must have a highly porous structure to maximize the surface area available
for efficient catalytic action. To summarize, the material properties required for each layer are
grouped together. For the membrane, the Nafion material must be uniformly hydrated and kept
below 100oC. When properly hydrated, the membrane is a good proton conductor and an
electronic insulator. The catalyst layers must be highly porous with an even distribution of
platinum catalyst. They must allow gas and water transport and be highly conductive both
thermally and electronically. The GDLs must also allow gas and water transport and be highly
conductive thermally and electronically. Last, the electrode plates must be mechanically strong to
support the MEA structure. They must be electronically conductive to deliver electrical power out
of the MEA. The electrode plates contain channels for reactant gas flow which not only delivers
the reactant gasses to the MEA, but also provides a location to humidify the input hydrogen gas
or evaporate water to the air flow. The electrode plates must be thermally conductive to dissipate
To construct the fuel cell, the MEAs are stacked so that they are effectively connected in
series. The anode plate of one cell is connected to the cathode plate of the next cell to create a
30
series electrical connection. For example, to construct a fuel cell with a 12 V output, if each cell
produces a voltage of 0.8 V, then 15 MEAs must be used to produce 12 V across the stack. Since
the cell voltage (MEA voltage) is a low value (e.g. 0.8V) and the cell current is high (e.g. > 3 A),
even a small contact resistance between cells will cause a significant loss in fuel cell stack
voltage. To minimize voltage loss at the connections between MEAs, the electrode plates are
combined. The anode plate of one cell is combined with the cathode plate of the next cell into one
physical plate called a bipolar plate. The use of bipolar plates over separate anode and cathode
plates is common and significantly reduces voltage losses of the fuel cell. The complication of
designing bipolar plates is the balance between minimizing electrical resistance between cells
and maximizing gas flow for the cells. Each bipolar plate must deliver hydrogen gas on its anode
side and deliver air on its cathode side. The channels for gas flow are thus formed on both sides
of the bipolar plate. There is a tradeoff between channel size and minimizing electrical resistance.
Striking a balance between these two concerns is one challenge of bipolar plate design. A second
pair of concerns is the thickness of the bipolar plate and the mechanical strength of the plate.
Minimizing the thickness of the bipolar plate minimizes the electrical resistance and weight of the
fuel cell but it reduces the mechanical strength. Traditionally, Poco graphite has been the
preferred material for bipolar plate construction. Properties such as low electrical contact
resistance, good mechanical strength, low weight, and the ability to withstand a corrosive
environment, are advantageous for bipolar plate construction. The disadvantages of using Poco
graphite are the cost, the brittleness of the material, and the difficulty of machining. For very thin
bipolar plates made from Poco graphite, mechanical stress and shock can crack the material.
High material cost and expensive machining processes contribute to higher fuel cell prices. Using
metal as an alternative to Poco graphite has been investigated and reported in literature.
Although metals are less expensive, less prone to crack under stress and shock, and are easier
to machine, they tend to corrode which causes higher resistances at contact points (Tawfik,
2006).
A practical PEMFC system consists of more than the fuel cell stack alone; it also includes
the BOP equipment and the hydrogen storage system. Using a commercial example of the
31
Horizon Fuel Cell Technologies H-100 fuel cell and a compressed hydrogen storage tank, figure
The H-100 fuel cell stack consists of 20 MEA cells and has a nominal operating point of 12 V at
8.3 A. At the nominal operating point, the cell voltage is 0.6 V. The maximum stack temperature is
65oC and the fuel cell stack is rated to operate between 5oC and 30oC ambient temperature
(Horizon, 2013). The BOP equipment that ships with the H-100 fuel cell include the fuel cell
controller, a fan on the fuel cell stack, a hydrogen supply valve, and a hydrogen purge valve. For
the H-100 fuel cell, the water management is self-contained inside the fuel cell stack (Horizon,
2013). The fuel cell controller gets input data from the temperature sensor and controls the fan to
regulate fuel cell temperature; the thermal management system. The fuel cell controller also
controls the hydrogen supply and purge valves. For the hydrogen storage system, the storage
tank contains compressed hydrogen at pressures up to 310 bar. The hydrogen pressure is
reduced in two stages. The high pressure regulator is included with the hydrogen storage tank
and steps down the hydrogen pressure from 310 bar down to 27.6 bar. The second stage, or low
pressure regulator, steps the hydrogen pressure down from 27.6 bar to 0.5 bar. The low pressure
regulator is purchased separately. Note that the fuel cell controller needs an external power
supply.
32
2.7 Proton Exchange Membrane Fuel Cell Polarization Curve
To characterize the performance of a fuel cell, a graph of the stack voltage versus the
fuel cell current can be used. Known as a polarization curve, the graph shows the operating
voltage over the full range of fuel cell operating current. The polarization curve can be measured
empirically by measuring the stack voltage as a connected electrical load is varied. The first point
to measure is the open circuit voltage; that is the stack voltage present when the electrical load is
disconnected. The electrical load is then connected and decreased from 10 mega-ohms down to
the lowest value (maximum electrical load) where the stack voltage is close to zero. Fuel cell
theory describes the features of the polarization curve from first principles. The theoretical model
can produce the polarization curve of a PEMFC design by adjusting the parameters of an
analytical model. The theoretical model can be thought of as an equation with four contributing
terms, or it can be represented as an electrical circuit. The four terms in the model are: the open
circuit voltage, activation loss, ohmic loss, and the concentration loss. In thermodynamic lingo,
the open circuit voltage represents the reversible voltage and the three loss terms represent
The open circuit voltage is a result of the chemical reactions taking place inside the fuel
cell. In other words, the change in chemical potential energy due to the change in chemical
configurations (anode and cathode reactions) result in an electrical voltage at the fuel cell
terminals. To quantify the change of chemical potential energy, the thermodynamic quantity
called the Gibbs free energy is commonly used. Gibbs free energy can be defined for a chemical
reaction as the energy available to do external work. The key point for determining the energy
released in a chemical reaction is that it is the change in chemical potential energy that is
available to perform external work. This is similar to a mechanical situation where a ball is at the
top of a hill. As the ball rolls down the hill, its velocity increases. The ball gains kinetic energy
which is equal to the change of potential energy. Another way of thinking about the chemical
reactions inside the fuel cell is that they change chemical potential energy into electrical potential
energy. For the overall chemical reaction inside the fuel cell, a hydrogen diatomic molecule is
33
combined with half of an oxygen diatomic molecule to yield a water molecule plus heat and
electrical energy.
1
H 2 O2 H 2O (8)
2
Analyzing this chemical reaction in more detail, the hydrogen, oxygen, and water molecules all
have a Gibbs free energy of formation. The difference between the Gibbs free energy of
formation of the product (the water molecule) and the Gibbs free energy of formation of the
reactants (the hydrogen and oxygen molecules) yields the chemical potential energy released
g f gf
H2O
gf
H2
1
2
gf
O2
(9)
Note that the lower case g with a bar above it denotes the Gibbs free energy per mole. Next,
1
Cathode reaction: 2e 2H O2 H 2O (11)
2
In these reactions, two electrons are released by the anode reaction and two electrons are
combined in the cathode reaction. Using conservation of energy, the chemical potential energy
gf n F E o (12)
g f
Eo (13)
n F
34
The minus sign in equation 13 accounts for the negative charge of the electron. The change in
Gibbs free energy is negative since energy is released from the chemical reaction, so the minus
signs cancel and give a positive fuel cell voltage. That is, the fuel cell cathode is positive relative
to the anode. The number of electrons participating in the chemical reaction must be accounted
for. This is similar to the mechanical analogy of two balls being placed at the top of a hill. Last,
there is the heat produced by the reaction. The amount of heat produced is the difference
between the change in enthalpy and the change in Gibbs free energy. Simply put, the change in
enthalpy is the total change of chemical potential energy, including any internal losses. The
change in Gibbs free energy is the total change of chemical potential energy minus any internal
losses. The internal losses are inherent to the chemical reaction and are thermodynamically
irreversible. The mechanical analogy to these internal chemical losses would be friction. If a ball
rolls down a hill, the kinetic energy gained is equal to the difference in potential energy minus
energy lost to friction. The energy lost to friction manifests as heat. Similarly, the chemical energy
lost to internal chemical losses manifest as heat. Additionally, the internal chemical losses also
determine the maximum theoretical efficiency of the fuel cell. The maximum efficiency a fuel cell
can have is simply the ratio of the change of Gibbs free energy to the change of enthalpy.
gf
max 100% (14)
hf
For example, at 60oC for the PEMFC reaction, the Gibbs free energy is -231.5 kJ/mole and the
enthalpy is -284.7 kJ/mole (Revankar & Majumdar, 2014). The maximum theoretical efficiency of
the PEMFC at 60oC is then 81.3%. The open circuit voltage also depends upon input gas
pressures and fuel cell temperature. Including these parameters yields the Nernst equation for
the PEMFC reaction. The Nernst potential is close to the true open circuit voltage of the fuel cell.
The actual open circuit voltage has to account for fuel crossover which is discussed at the end of
this section.
o
g f
Nernst equation: Eo
2F
RT
2F
ln PH2
RT
4F
ln PO2 (15)
35
where: E o = open circuit voltage of PEMFC without accounting for fuel crossover (V)
T = temperature (K)
At the electrodes, the chemical reactions have a finite rate of reaction. Due to the limited
reaction rate at the electrode surfaces, a portion of the cell voltage is used to drive the chemical
reactions. This energy loss is known as activation loss. It is the dominant energy loss mechanism
at low fuel cell currents. At zero fuel cell current, the activation loss is negligible. As fuel cell
current increases from zero, the activation loss increases logarithmically which give the
polarization curve a distinctive exponential decay for small fuel cell currents (Madakannan,
lecture notes 2016). Even when there is no fuel cell current, there are reactions taking place at
both electrodes. Transfer of electrons from the electrode to the electrolyte is balanced by a
transfer of electrons from the electrolyte to the electrode so that there is a net zero current across
the interface. At zero fuel cell current, the current density of just the forward reaction is known as
the exchange current density, io (mA/cm2). The magnitude of the exchange current varies,
depending upon the chemical activity at the electrode surface; typically ranging from 10-6 mA/cm2
to 102 mA/cm2. Factors such as electrode and electrolyte materials, electrode surface porosity,
catalyst activity, and temperature all contribute to the magnitude of the exchange current. Larger
values of the exchange current result in lower activation losses. For the PEMFC, the rate of
reaction is much faster at the anode than the cathode. The exchange current density is much
higher at the anode than at the cathode. The end result is that the activation loss at the cathode is
the dominant contributor the overall fuel cell activation loss. An equation for the activation
36
RT i
Vact ln (16)
2 F io
resistance of the electrodes and cell interconnects and the ionic resistance of the membrane,
ohmic losses drop an internal voltage equal to the product of total internal resistance and fuel cell
current. Ohmic loss is the dominant loss mechanism at midrange fuel cell currents.
Vohmic i R (17)
The third loss mechanism in PEMFCs is concentration loss. When current is drawn from
the fuel cell, the reactant gas concentrations at the electrode-electrolyte interface is lowered. As
fuel cell current increases, the reactant concentration at the electrode continues to decrease. At
low current densities, the depletion of reactant gas concentration is small and the voltage loss
due to concentration loss is negligible. However, when the fuel cell current is large, the gas
concentration depletion at the electrode surfaces become significant and the concentration loss
becomes a limiting factor. In other words, the reactant gasses have a maximum diffusion rate to
the electrode surfaces. When reactant gasses can no longer be supplied fast enough, the
reactant concentration at the electrodes drops to zero, which in turn causes the fuel cell voltage
to drop to zero. Concentration losses are only significant near the maximum fuel cell current and
are the limiting mechanism for maximum fuel cell current (Madakannan, lecture notes 2016).
37
J
Vcon B ln 1 (18)
Jmax
Combining the Nernst voltage and the three loss mechanisms gives the working voltage
The total output voltage of the fuel cell stack is then the sum of cell voltages present in the fuel
cell stack.
38
For the complete polarization curve shown in figure 12, there are three regions: the
activation region, the ohmic region, and the concentration region. In each region, the
corresponding loss mechanism is dominant. Note that the open circuit voltage is less than the
Nernst voltage. This is due to the last loss mechanism, fuel crossover. Due to the fact that the
membrane is not perfect, some fuel leaks over from the anode to the cathode and some oxygen
leaks over from the cathode to the anode. Also, to a small extent, some electrons leak through
the membrane as well. The result of fuel crossover is that recombination occurs which causes a
constant voltage loss across the entire polarization curve; usually around 0.2 volts (Larminie &
Dicks, 2003). The end result is that the actual open circuit voltage of a single cell is slightly less
than one volt and that the open circuit voltage for the entire fuel cell stack is the number of cells
times the cell voltage. For example, the Horizon H-100 PEMFC has 20 cells and an open circuit
voltage of 19 V (Horizon, 2013). The open circuit voltage of one cell is calculated to be 0.95 V.
material of the electrode and the electrolyte. On the electrode side, a thin layer of electrons
accumulates at the surface, while on the electrolyte side, a thin layer of cations accumulate near
the electrode surface. The net effect of the charge accumulation is that of a capacitor. This effect
is universal in solid electrodes immersed in liquid electrolyte systems. The first model of the
electric double layer was developed by H. Helmholtz in the 1850’s. In the Helmholtz model, the
negative and positive charge accumulation was modeled as a parallel plate capacitor. The
electron charge concentration at the electrode surface is analogous to the negative capacitor
plate. The cation concentration in the electrolyte near the electrode surface takes on the role of
the positive capacitor plate. The effective distance between the plates is the radius of one
solvated cation which results in a parallel plate capacitor model (Revankar & Majumdar, 2014).
A
C (21)
d
Due to the charge accumulation of the electric double layer, a voltage difference is present across
the interface and is the source of the activation loss discussed earlier. The electrode surface is
very porous and has an effective surface area roughly103 times greater than the simple geometric
surface area of the electrode. The distance between the charge layers is typically about 20
angstroms which results in an effective capacitance of several farads for a common PEMFC
(Larminie & Dicks, 2003). Since the Helmholtz model, more refined models of the electric double
layer have been developed such as the Gouy-Chapman model in 1910, the Stern model in 1924,
and the Grahame model in 1951 (Revankar & Majumdar, 2014). The capacitance due to the
electric double layer affects the dynamics of the fuel cell response to sudden changes in current.
The voltage has a delayed response to a step increase or decrease in current load. As the
current load changes abruptly, the voltage response comes to its equilibrium value in a smooth
curve-like fashion, similar to the voltage response of an RC circuit. An electric circuit model of the
double layer phenomena places a capacitor in parallel with a resistance that represents the
An important and challenging topic for hydrogen powered and hybrid powered UAVs is
hydrogen storage. Currently, there are four methods of hydrogen storage: compressed gas, liquid
hydrogen, solid state chemical, and physical absorption (Zhang, 2014). The compressed gas and
liquid hydrogen methods are well developed technologies and are used in many hydrogen
powered UAV applications. For example, the NRL Ion Tiger has used compressed hydrogen for
its 2009 demonstration flight and liquid hydrogen in its 2013 flight (Swider-Lyons, 2011, “Ion Tiger
Fuel Cell Powered UAV”, 2016). Although compressed gas and liquid hydrogen are considered
mature hydrogen storage technologies, there is continuing innovation to increase the volumetric
and gravimetric density. To use target metrics from the automobile industry, the Department of
Energy (DOE) has published design goals for hydrogen storage in light duty vehicles. In the
Freedom Car Fuel Partnership report published in 2009, the DOE lists 2015 target values for the
40
volumetric density and gravimetric density as 40 g/L and 5.5% respectively (“Targets for Onboard
Hydrogen Storage Systems for Light-Duty Vehicles”, 2009). The gravimetric density is another
way of stating the mass density, but includes the weight of the storage tank. Gravimetric density
is useful in that it measures the effectiveness of the storage system. For example, if a
compressed gas tank can store hydrogen at a very high mass density, but has an enormous
weight, then it will not be very useful for vehicle applications. Gravimetric density takes the
storage tank weight into account by stating the ratio of hydrogen weight to the storage tank
mH2
Gravimetric density: g 100% (22)
mtan k mH2
There is a wide variety of compressed gas tanks used in experimental UAV applications.
Typically, low end UAV compressed gas storage tanks are compressed air tanks from the
paintball industry. For example, the University of Johannesburg Piper Cub used a 0.255 kg CO2
paintball cylinder to store hydrogen. Although the cylinder was capable of holding 220 bar of
pressure, the Piper Cub team only filled it to 30 bar (Furrutter, 2009). This may have been done
for safety reasons. The aluminum cylinder was designed to store CO2 and not hydrogen. If the
hydrogen gas diffuses into the aluminum cylinder walls, it can weaken the aluminum cylinder
causes a loss of metal ductility similar to stress corrosion cracking. The result is that the energy
required to form stress cracks is reduced and the strength of the aluminum cylinder is derated
(Madakannan, lecture notes, 2015). In some commercial storage tanks, a polymer liner is used to
combat hydrogen embrittlement (Zhang, 2014). In contrast to the paintball storage tank, high end
compressed gas storage tanks have been developed for UAV applications. For the Ion Tiger
project, the NRL team used a custom aluminum tank with a carbon overwrap. The carbon
composite overwrap reinforces the aluminum tank. The carbon overwrapped aluminum hydrogen
tank weighed 3.6 kg with a water volume of 22 L and had the capacity to store 500 g of hydrogen
(Swider-Lyons, 2016). Using the ideal gas equation, the calculated pressure of the tank is found
41
to be 280 bar (about 4000 psi). There is a discrepancy though, the reported pressure of the tank
is 5000 psi (Swider-Lyons, 2016). When highly compressed, hydrogen can no longer be treated
as an ideal gas. At high compressions, intermolecular forces can no longer be ignored and a
P V n Z R T (23)
n = number of moles
T = temperature (oK)
Taking the ratio of 5000 psi over 4000 psi, the compressibility factor must be 1.2. Referring to the
compressibility factor chart, a compressibility factor of 1.2 is a valid figure for 4000 psi at 300oK
(Zhang, 2014). The performance of the carbon overwrapped aluminum storage tank used by the
NRL team is excellent, coming in at a volumetric density of 22.7 g/L and a gravimetric density of
12.2%. Compared to the DOE 2015 targets, the NRL storage tank falls short of the 40 g/L
volumetric density goal but exceeds the 5.5% gravimetric density goal. Next, a commercially
available compressed air paintball storage tank, the Ninja ProTM 90/4500, is evaluated. Made out
of carbon fiber, the 90 cubic inch compressed air tank can hold up to 4500 psi (310 bar) of
pressure. If used for hydrogen storage, and including a 20% safety factor (pressurizing to only
248 bar), the Ninja Pro 90/4500 tank can hold 24.6 g of hydrogen. With a weight of 1.45 kg and a
water volume of 3.56 L, the volumetric density works out to be 6.9 g/L and the gravimetric density
is 1.7% (“Ninja Pro v2 90/4500 specs”, 2016). Considering the output flow of hydrogen at
standard temperature and pressure (STP), the Ninja Pro v2 90/4500 tank can deliver 298.9 liters
of hydrogen. If the Ninja Pro v2 90/4500 tank is delivering hydrogen to a Horizon H-100 PEMFC
(feed rate of 1.3 SLPM), it can keep the fuel cell running at full power (100 W) for 3 hours and 50
42
minutes. In other terms, the Ninja Pro v2 90/4500 supplying a Horizon H-100 fuel cell, can store
383.2 Wh of energy.
Figure 13. Ninja Pro v2 90/4500 Compressed Air Tank (www.ninjapaintball.com, 2016)
Typically, APM system design for UAV applications use DC-to-DC converters. One or
more power sources are connected to the electrical bus through DC-to-DC converters. The
function of the DC-to-DC converter is essentially that of a voltage regulator. The input voltage to
the DC-to-DC converter may vary over a wide range while the converter maintains the output
voltage at a fixed setpoint regardless of the load current. With many off-the-shelf DC-to-DC
converters, the output voltage is fixed and cannot be changed. For most of the DC-to-DC
converters used in APM systems, the output voltage is not fixed and can be controlled by a
microprocessor. A large majority of the APM systems discussed in UAV APM research literature
use one or more DC-to-DC converters in the main power path to control power flow. The voltage
sent to the power bus from one or more sources is controlled by a DC-to-DC converter which
determines the power split. Figure 14 shown below illustrates three circuit concepts that use DC-
to-DC converters.
43
Figure 14. APM Circuit Concepts using DC-to-DC Converters
In part (a) of figure 14, the fuel cell is connected to the bus through a DC-to-DC converter while
the battery is connected directly. In this case, the battery voltage determines the bus voltage. The
fuel cell voltage is varied just below the bus voltage to control the power split. If the fuel cell
voltage is adjusted above the battery voltage, then the fuel cell supplies 100% of the ESC load
and the battery may recharge. Depending upon the battery SOC, the bus voltage may be less
than the fully charged battery voltage. The microprocessor must monitor the bus voltage to
accurately control the power split. In part (b), the fuel cell is connected directly to the bus and the
battery is connected through a DC-to-DC converter. In this case, the fuel cell voltage determines
the bus voltage. Here, the battery voltage is adjusted to control the power split to the ESC load.
The battery voltage is adjusted just above the bus voltage to determine the power split. If the
battery voltage is set below the bus voltage, then the fuel cell delivers 100% of the load power
and may charge the battery. Note that the DC-to-DC converter is two-way; that is power can
desired. As in the circuit presented in figure 14, part (a), a supervisory microprocessor must
monitor the bus voltage in order to accurately control the power split. In part (c), both fuel cell and
battery are connected to the bus through DC-to-DC converters. This is the most flexible scenario
44
and has the advantage that the bus voltage can be maintained at an optimum value for the ESC
load. In this scenario, the relative values of the fuel cell and battery voltage will determine the
power split to the ESC load. This third scenario is unique in that two DC-to-DC converters are
connected directly to each other through the bus. There are several methods to control the power
split between the two converters, namely the droop-share method or the driver booster method
(“Current Sharing in Power Arrays”, n.d.). In the droop share method, the output impedance of
one converter is varied to control the power split. An error signal is used in the DC-to-DC
converter control loop to cause the output voltage to be a function of both the setpoint and the
converter output current. In the driver booster method, one converter acts as a driver and the
second converter acts as a booster. The driver sets the bus voltage while the booster augments
bus current when needed. In the driver booster method, microprocessor control coordinates the
action of both converters (“Current Sharing in Power Arrays”, n.d.). In general, there are several
disadvantages to using DC-to-DC converters in UAV APM systems. First and most important is
the weight of the DC-to-DC converter. As the power rating of a DC-to-DC converter increases, so
does its weight. For example, a commercially available 3 W DC-to-DC converter weighs around 2
g while a 400 W converter weighs around 220 g. When used for a low power load, such as
supplying a microprocessor circuit, the weight of the DC-to-DC converter has little impact on UAV
weight. However, when used to convert the output voltage of the fuel cell or battery (main power
path), then the DC-to-DC converter weight has a detrimental impact on overall UAV weight. Note
that using a DC-to-DC converter on the battery output is more costly in weight than using one on
the fuel cell output. The DC-to-DC converter power rating must match the maximum power
expected. With the battery supplying higher power than the fuel cell, the converter weight will
naturally be higher. The secondary disadvantages of DC-to-DC converters are cost, size,
reliability, and complexity. Third, DC-to-DC converters radiate electromagnetic interference (EMI)
which may interfere with the radio control receiver. Last, DC-to-DC converters are efficient in
converting power; typical power conversion efficiencies range from 82% to 95%, but they are not
45
Fuel cells operate best under constant load conditions. Testing fuel cell performance
under pulsed load conditions reveals limitations. Consider a pulsed current load where the low
and high load currents are 10% and 90% of the fuel capacity respectively. At the low load current
level, excessive air flow across the cathode can cause the membrane to dry out. As a result, the
membrane ionic resistance increases. When the high current pulse occurs, the stack voltage now
dips lower due to increased internal resistance. Due to increased power dissipation across the
membrane, the membrane temperature rises (Jones, Mepsted, Moore, 1998). The result is a
repeating cycle of membrane dehydration and excessive membrane temperatures. Also, during
the high load current pulses, there is a limit to the change in rate for which the air can diffuse to
the cathode. In effect, the abrupt increase in current load causes oxidant starvation at the
cathode (Suh, 2006). Lower oxidant partial pressure due to starvation then drives the stack
voltage dip further down. Pulsed current operation tends to cause excess water build up at the
cathode, which in turn, tends to block oxidant diffusion (Suh, 2006). The result is a cycle of
oxidant starvation and excess water build up at the cathode. Purging of excess water is required
To protect the fuel cell against frequent transient loads, a battery can be combined with
the fuel cell to create a hybrid power system. The battery can be connected directly across the
fuel cell as in the PPM case, or it can be connected to the fuel cell through a DC-to-DC converter
as in the APM case. A study conducted in 1998 compared the pulsed load performance of a
PEMFC, a lead acid battery, and the two sources combined (Jones, Mepsted, Moore, 1998). The
fuel cell alone was able to drive the load, but excessive voltage drop of the fuel cell stack was
problematic and frequent purging of the fuel cell was necessary. The lead acid battery was able
to power the load, but it could only drive the load for 3 hours and 20 minutes. When the two
sources were connected in a PPM configuration, the load was driven for six days continuously
and no fuel cell purging was necessary. The key difference of the hybrid performance over the
fuel cell alone was that the battery charged during low load current intervals. The battery charge
current kept the fuel cell current above a minimal value which was enough to keep the membrane
hydrated and to prevent water buildup at the cathode (Jones, Mepsted, Moore, 1998).
46
One difficulty with the PPM configuration is that the VI-curves of the battery and fuel cell
must be closely matched. If the fuel cell VI-curve is of a much higher voltage than the battery VI-
curve, then the fuel cell current capacity may limit the hybrid maximum current and the battery will
tend to overcharge. If the fuel cell VI-curve is of a much lower voltage than the battery VI-curve,
then the fuel cell will consume battery power and the PPM system will be ineffective. Using an
APM configuration allows the fuel cell VI-curve to differ from the battery VI-curve. Using a DC-to-
DC converter between the fuel cell and battery lifts the restriction of matching VI-curves and
delivers higher peak currents during pulsed load operation. A study performed in 2003, examined
APM hybrid power system performance under pulsed load conditions (Jiang, Blackwelder,
Dougal, 2003). Three control strategies were used to drive the DC-to-DC converter. First, the
maximum power strategy, operated the fuel cell at the maximum power point. Under the
maximum power strategy, a battery starting at a SOC around 70% would recharge and remain at
100% SOC. Second, the maximum efficiency strategy; the fuel cell operating point was set to
around 70% of maximum current. The battery SOC would slowly decrease until completely
discharged. Third, the adaptive strategy, was a combination of the first two strategies so that the
battery could be maintained within a SOC target range. When the battery SOC was below the
upper threshold, maximum power strategy was active which slowly recharged the battery. When
the battery SOC reached the upper threshold, maximum efficiency mode became active which
allowed the battery to slowly discharge. When the battery SOC reached the lower threshold
value, the DC-to-DC controller then switched back to maximum power mode. The maximum
power, maximum efficiency cycle would then repeat to keep the battery SOC between the upper
A variety of APM control algorithms are presented in literature such as the study
completed in 2007 by Jiang et al. (Jiang, Gao, Dougal, 2007) and the study completed in 2014 by
Lee and Park (Lee & Park, 2014). In the first paper by Jiang et al., the amount of fuel cell current
delivered to the bus is a piece-wise linear relationship that depends upon the battery voltage
which is used as a measure of the battery SOC. Similar to the 2003 study by Jiang et al. (Jiang,
Blackwelder, Dougal, 2003), the fuel cell current setpoint is set at either the maximum power
47
point or the maximum efficiency point. However, in the Jiang et al. 2007 study, instead of
switching between maximum power mode and maximum efficiency mode only, a new hybrid
mode for the DC-to-DC converter combines the two modes. Between the maximum and minimum
battery voltages, a linear combination of the two modes are used which connects the two
operating points by a line. In the 2014 study by Lee and Park, solar cells are added as a third
power source and are also connected to the electrical bus through a DC-to-DC converter. The
control algorithm uses battery SOC and load power to create 6 sectors of operation where each
sector defines a predetermined power split between the three power sources.
An APM electrical architecture that does not use DC-to-DC converters was developed at
ASU Polytechnic campus for the Composite UAV project (Monaco, 2013). Although the APM
system was only tested in the lab and not used in actual flight, it serves as an example of an APM
approach that is further developed in this research. In the ASU APM circuit, the fuel cell is directly
connected to the bus while the battery is connected to the bus through a switch. When the switch
is off, then only the fuel cell powers the load. When the switch is closed, then both battery and
fuel cell deliver power to the load. During the takeoff and initial climb phases of the flight profile,
the switch is closed; maximum power is delivered to the load. When the cruise flight phase is
reached, the switch is then opened and only the fuel cell powers the load. The salient feature of
this approach is that the high energy losses of the PPM method at low load currents is avoided
while both power sources are used at high load currents. Since the energy losses of a PPM
circuit are low when the load currents are high, the ASU APM circuit reaps the benefits of a PPM
architecture while avoiding the power contention issue at low load currents. The ASU APM circuit
complexity is low; only the switch and a method to control the switch is needed. The cost and
weight are low since only a few electronic components are needed. The benefits of an APM
approach are combined with the benefits of the PPM approach while avoiding the excessive
energy losses of PPM architecture at low load currents. The actual electronic components used
to construct the ASU APM circuit did not use a switch, but instead used a metal oxide field effect
transistor (MOSFET) to perform the function of the switch. The MOSFET was then controlled by
48
an Arduino microprocessor (Monaco, 2013). During takeoff and initial climb phases of the flight
profile, the microprocessor controls a bias network (a circuit consisting of resistors to operate the
MOSFET) so that the MOSFET is in its saturation region of operation with the controlling gate-to-
source voltage (Vgs) at its saturation value. This minimizes the drain-to-source resistance (rds) of
the MOSFET and is analogous to the switch being closed. Typical values of rds for a power
MOSFET in the on-state are in the 2 mΩ to 100 mΩ range (rds = rds(on)). When the cruise phase of
the flight profile is reached, the microprocessor then turns off the MOSFET by changing the logic
state of its controlling pin to the bias network. The MOSFET is then in an off-state where rds > 10
MΩ (rds = rds(off)); this is analogous to the switch being open. This example demonstrates that an
effective APM system can be constructed from a simple, low weight, low cost, electronic circuit.
During lab testing of the ASU APM circuit, the MOSFET transistor burned up and the
circuit was abandoned in favor of another approach using a relays. The ASU Relay circuit
performed as expected and was used for further testing. The failure of the MOSFET was
attributed to the current rating. Under test conditions, the battery current was measured at 50 A
while the MOSFET was rated for only 30 A (Monaco, 2013). Even though the MOSFET had a
current rating that was too small for the actual current conditions of the circuit, a MOSFET with a
current rating of 50 A or larger may have failed as well. The reason is that proper heat sinking is
also required to operate a power MOSFET under high current conditions. In other words, if a 60 A
power MOSFET was used instead, but had no heat sink, a 50 A current load would have caused
the MOSFET to fail anyway. In typical power MOSFET data sheets, maximum junction
temperature is listed (usually around 150oC) along with thermal resistance ratings such as
junction-to-ambient and junction-to-case. Circuit design must ensure that the MOSFET junction
temperature is kept below the maximum allowable value. Although the ASU APM circuit did not
succeed for its intended purpose, there are several features of the circuit that suggest a different
approach to an APM design that is discussed further in section 4.1. Shown below in figure 15 is a
49
Figure 15. ASU APM Circuit (Monaco, 2013)
In figure 15, the ASU APM circuit shows the microprocessor controlling the state of Q1
which connects or disconnects the battery from the bus. The points A and B represent the
microprocessor monitoring the fuel cell and battery voltage levels. Diodes D1 and D2 prevent any
reverse current flows between the power sources. The original design intent was to either fully
connect or fully disconnect the battery from the bus. In this scenario, the diodes are necessary
since the battery is effectively connected directly across the fuel cell. When Q1 is in the on-state,
the circuit is similar to the PPM case where power contention may occur between the sources.
The diodes prevent power exchange between the sources. There is a drawback to the diodes
though, they drop a forward bias voltage and thus dissipate power during normal operation. The
power dissipation of the diodes are significant since they are in the main power path.
50
CHAPTER 3
SIMULATION COMPONENTS
In this study, system simulations are performed in the MatlabTM environment from
Mathworks, Inc. Within the Matlab SimulinkTM environment, SimscapeTM physical modeling is
used to develop power system components which are then assembled into a hybrid power
system model. Simscape is designed to model physical systems using a Physical Network
Approach (PNA) (“Simscape User’s Guide”, 2014). Simscape blocks are placed in a Simulink
model to represent a system component, however Simscape blocks are fundamentally different
from Simulink blocks. Simulink blocks represent mathematical operations while Simscape blocks
represent physical components. Assembling Simulink blocks into a model results in a graphical
information called signals and define an order of mathematical operations defined within Simulink
blocks. Simulink blocks are considered causal since for a given input, an appropriate output is
developed. Simulink signals always have a direction associated with them and do not carry units.
On the other hand, assembling Simscape blocks results in a graphical model of a physical
system. Connections between Simscape blocks represent physical connections between physical
components and do not define any order of mathematical operations. Simscape blocks have two
types of ports, physical ports and physical signal ports. Physical ports carry a flow of energy in or
out of a block and do not have a direction. When physical ports on two Simscape blocks are
connected, energy can flow between the two components in either direction. Simscape models
are considered non-causal since the blocks have no input or output; models consist of energy
flow between blocks. Connections between physical ports always carry units. Physical signal
ports carry information about a physical value (with units) and do have a direction assigned.
Physical signal ports are not part of the energy flow network of a physical model; instead, they
can be used to monitor internal block variables or carry information between blocks.
Within a Simulink model, the canvas is the window working and graphical area where
blocks are placed, connections are made, and the graphical presentation of the simulation model
51
is displayed. On the canvas, both Simulink blocks and Simscape blocks are placed, but they are
still functionally separated. Simulink blocks inherently work inside the Simulink canvas without
any auxiliary helper blocks. Simscape blocks require a solver block to be connected to the
reference node of the Simscape model. One great feature of the Simulink environment is that
Simulink and Simscape models may be connected together on the same canvas. Observer and
translation blocks are used to read and convert a Simscape physical value to the Simulink model
domain. A second great feature of the Simulink environment is the ability to use Matlab scripting
files to automate simulations. Since Simscape models are contained within the Simulink canvas,
both Simulink and Simscape models, or combinations of the two, can be automated through
Matlab scripts. The result is that a physical model may be simulated for a range of parameter
values and the results from each simulation run can be extracted, grouped, and graphed to reveal
The Simscape physical modeling method allows for the creation of custom blocks
through the use of the Simscape text-based language. A custom system component can be
described by a text file and then compiled into a Simscape library. Once compiled, the custom
component has a graphical block inside the library that may be dragged onto the Simulink canvas
like any other block. Although the Simscape language is similar to the Matlab language, it is
different and has a required file structure to create a Simscape component. A Simscape
component file has three main sections: a declaration section, a setup section, and an equation
section. In the declaration section, through and across variables, physical ports, physical signal
ports, and parameters are defined. In the setup section, the relationships between the through
and across variables are defined. The setup section also defines initial conditions and provides a
location for parameter checking functions. The third section, the equation section, defines the
For a hybrid power system simulation, the battery model employed must strike a balance
between fidelity and computational efficiency. A good lithium ion battery model for the system
level simulation purposes of this study is presented by Olivier Tremblay and Louis-A. Dessaint
52
(Tremblay & Dessaint, 2009). The model equations take battery capacity, SOC, and current
delivery history into account. In this model, discharge and charge equations are separate. The
equation for the discharge case (i > 0) is listed below in equation 24.
Vbat Eo i R A exp B i (t )dt Kc
Q
Q i (t )dt
i (t )dt i * (24)
The equation for the charge case (i < 0) is listed below in equation 25.
Vbat Eo i R A exp B i (t )dt Kc
Q
Q i (t )dt
i (t )dt KR
Q
i* (25)
i (t )dt 0.1 Q
where: Vbat = battery voltage (V)
The Tremblay battery model is based from earlier work by Shepherd (Shepherd, 1965). The key
difference is the use of filtered current i*. The filtered current is the instantaneous battery current
after it has been passed through a low pass filter. When implementing the Tremblay model in
Simulink, the filtered current prevents an algebraic loop (Tremblay & Dessaint, 2009). In this
study, the Tremblay model is implemented using the Simscape language and the algebraic loop
is no longer a problem. In this Simscape model, battery current can be used in place of the
53
filtered current. The Simscape code to implement the lithium ion battery model is listed in
appendix A.
As a first test of the battery model, a commercially available LiPo 4S 8 Ah, battery is used
as an example. To plot battery discharge curves for a range of discharge rates, a Simscape
model is developed which loads the battery at a regulated current level. The load circuit used is a
constant current regulator circuit with a programmable current setpoint (figure 16).
Transistor Q1 (p-channel MOSFET) is the regulating element which determines the load current
drawn from the battery under test. Resistor R2 functions as a current sense resistor and is fixed
at a low resistance value of 0.1 Ω. Operational Amplifier (OpAmp) U1 measures the differential
voltage across R2 and amplifies this voltage by a factor of 20. OpAmp U2 then takes the
difference between the current setpoint voltage and U1 output to drive the gate voltage of Q2.
Transistor Q2 (n-channel MOSFET) in conjunction with resistor R1 then control the gate voltage
of Q1. As the battery voltage changes, the current drawn from the battery is held at a fixed level
The current setpoint voltage is a linear relation to the resulting load current on the battery. The
first term is a 1.7 V offset which sets the quiescent operating point of the gate voltage of Q2. The
second term relates the desired load current on the battery to the Q2 gate voltage rise above the
54
quiescent value. The end result is a programmable constant current load circuit that can be used
The battery model parameters are set to represent a typical LiPo battery pack used by
available (“MaxAmps LiPo 4S 8Ah”, 2016). This particular battery is in the midrange of the
product family and is a good compromise between capacity and weight. The size, weight, and
electrical capacity are well matched to the ASU Composite UAV and will be used as the reference
battery in the power system simulations of chapter 4. With a nominal voltage of 14.8 volts and
battery pack weight of 0.764 kg, the specific energy density of this 8 Ah battery pack works out to
be 155 Wh/kg. A nice feature of this battery pack is the high discharge rating. Using two 4S, 4 Ah
lithium ion batteries in parallel, the maximum discharge rating is reported to be 150 C. Operating
at half the maximum discharge rate (75 C), the battery pack can deliver 600 A of current for 48
seconds yielding a specific power density of 11,623 W/kg. The Tremblay battery model
parameters used to represent the MaxAmps 4S 8Ah battery pack are given in table 3.
Table 3
The discharge curves for the LiPo 4S battery are obtained by Simscape simulations of the
discharge circuit (figure 16) over a range of load currents. For each simulation run, battery
voltage and delivered charge data is extracted. The data is then combined and graphed together
to yield the plot in figure 17. The Matlab automation script is listed in appendix B.
55
Figure 17. LiPo 4S 8 Ah Battery Discharge Curves
To compare the LiPo 4S battery model against a Simscape battery model from the Sim Power
Systems library, a second discharge simulation is created. It compares the battery voltages over
a 2 hour period when both are supplying a 4 A load (0.5 C discharge rate). The Simulink model is
listed in appendix C and the graph of battery voltages versus time is shown below in figure 18.
56
3.3 Battery Charger Model
To recharge the LiPo 4S battery during flight, an onboard battery charger circuit is
included with the APM system. A standard lithium ion battery recharge cycle consists of a
constant current phase followed by a constant voltage phase (Tsang, Sun, Chan, 2010). The
constant current phase is used during most of the charge cycle, and then the constant voltage
phase is used at the end of the cycle to top off the battery charge. To simplify the charger circuit,
only the constant current phase is implemented. The battery charger circuit (figure 19) uses a
small DC-to-DC converter to generate a 24 V source and then uses a programmable constant
Input power from the electrical bus is fed to the DC-to-DC converter which produces a constant
control the current flow to the battery. Similar to the constant current load circuit described in
section 3.2, the difference between the setpoint voltage and the sense resistor voltage is sent to
the gate of Q2 (n-channel MOSFET) which in turn, controls the gate voltage of Q1. The key
difference here is the resistor R3, which drops the proper amount of voltage for a specified
recharge rate. The value of R3 will depend on the range of recharge rates desired. For this case,
the recharge circuit is designed to operate near a 0.1 C recharge rate which sets the value of R3
57
to 8.25 Ω. A Simscape model is created to test the recharge circuit on the LiPo 4S battery (figure
20). Note that the recharge rate will determine the size of the DC-to-DC converter necessary. For
this circuit, the 24V output of the converter delivers 0.8 A which allows for a 20 W converter to be
used. Minimizing the DC-to-DC converter power rating keeps the recharge circuit weight to a
minimum.
The simulation starts out with the battery at a 40% battery SOC and is run for 5.5 hours at a
charge rate of 0.1 C (0.8 A). At the end of the charge cycle, the battery is at 96% SOC (figure 21).
The corresponding battery voltage and recharge current are shown in figures 22 and 23
respectively.
58
Figure 22. Battery Voltage during Recharge
The PEMFC has a variety of mathematical models to choose from. The type of simulation
will determine which model is appropriate. For purposes of this study, an electrical circuit model
has a good balance between model fidelity and computational cost. The electric circuit equivalent
model for a PEMFC consists of four sections; an ideal voltage source and three series resistors.
In thermodynamic lingo, the ideal voltage source represents the reversible Nernst voltage while
the three resistors represent irreversible losses. In electrochemical language, the three sources of
energy loss are called polarizations and consist of the activation loss, the ohmic loss and the
concentration loss. To improve the fuel cell model, the effect of the electrochemical double layer
can be included. A capacitor added across the activation and concentration resistors causes a
59
delay between fuel cell voltage and current. Shown below in figure 24, part (a) shows the electric
circuit equivalent model of a PEMFC without the electric double layer capacitor. The fuel cell
voltage response to a step increase in current has no delay. An improved model includes the
Note that the electric circuit may represent a single cell or the entire fuel cell stack. The values
assigned to the electrical components are either for a single cell or multiplied by the number of
cells in the stack. In standard electrical circuit analysis, resistors are assumed to be fixed in value.
In the PEMFC equivalent circuit, the resistor values are current dependent which makes the
The Simscape component for the PEMFC electrical circuit model has several ways it can
be implemented. The entire circuit can be implemented at a single level (flat design) using
standard Simscape blocks connected graphically, or it can be implemented in a single text file.
The difficulty with this approach is that a flat design results in a very complicated file, graphically
or text-based. To break the problem down into simpler pieces, each electrical component within
the circuit can have its own file. The sub-component files can then be assembled into the
electrical circuit (hierarchical design). The top level file connects the sub-components and can be
60
graphical or text-based. In this study, the PEMFC Simscape component implements the electrical
circuit model by a hierarchical approach using text-based files for the sub-components and the
top level.
have been reported. Buasri and Salameh suggest an electrical model similar to figure 24 (b), but
add a series diode between Ract and Rcon (Buasri & Salameh, 2006). The exponential region of
the diode forward conduction mode is similar to the exponential decay of the fuel cell polarization
curve in the activation region (see figure 12). In the Buasri model, expressions for Vact and
Vohmic are unchanged from the standard expressions given in equations 16 and 17 respectively.
However, the Buasri model suggests a new expression for V con which is quite different from the
standard equation given by equation 18. In equation 27 shown below, the Buasri concentration
loss expression has two parameters, K1 and K2 which are determined empirically (Buasri &
Salameh, 2006).
V con K 1 i FC e K 2 i FC (27)
A second approach to the PEMFC electrical circuit model uses the same circuit as shown
in figure 24 (b), except the expressions for Eo, Vact , and Vohmic have been modified from the
standard expressions given by equations 15, 16, and 17 respectively (Lee & Wang, 2007,
Qingshan et al., 2008). In this model, referred to as the Lee model, the expression for V con
remains unchanged from equation 18. In this study, the Lee model is compared to the standard
expressions given in section 2.7 and is used as the PEMFC electrical model for the hybrid power
system simulations of chapter 4. The Simscape PEMFC component developed in this paper is an
To calculate the Nernst potential using the standard method, the Nernst equation is used
o
g f
Nernst equation: E o
s tan dard
2F
RT
2F
ln PH 2
RT
4F
ln PO2 (28)
61
Table 4
Gibb’s Free Energy for Hydrogen Reaction (Madakannan lecture notes, 2016)
Form of Water Temperature
Product (C) g (kJ/mol)
liquid 25 -237.2
liquid 80 -228.2
gas 80 -226.1
gas 100 -225.2
gas 200 -220.4
In the Lee model, the Nernst potential is approximated by equation 29 (Lee & Wang, 2007). The
o
ELee
1.229 0.85 103 T 298.15 4.3085 105 T ln PH2 0.5 ln PO2
(29)
In the Lee model, the approximation for the Nernst potential does not require a value for the
Gibb’s free energy, only temperature and gas pressures are needed. The percentage error using
the Lee Nernst potential expression over temperature and hydrogen pressure can be illustrated
by a surface plot of the percentage difference (Matlab code listed in appendix E).
E0 ELee
o
Eo s tandard
o 100% (30)
Es tandard
62
Figure 25. Percent Error of Lee Model Approximation of Nernst Potential
From figure 25, the Lee model is within +4% and - 3% of the standard Nernst potential
calculations. The percent error is predominantly dependent upon hydrogen pressure and not
temperature. The green grid is the zero-error plane which shows minimum error when the
hydrogen pressure is near 2 atmospheres. In figure 26, a different view of the same data is
helpful to determine the percent error as a function of hydrogen pressure. Each line is the percent
63
Figure 26. Percent Error of Lee Model for Different Hydrogen Pressures
To calculate the activation polarization using standard methods, a value for the charge
transfer coefficient and a value for the exchange current density i o are required.
RT i
Vact s tandard ln FC (31)
2 F io
The expression for the activation loss used in the Lee model is shown below in equation 32 and
only requires values for temperature and air pressure (Lee & Wang, 2007). Note that PO is the 2
Vact Lee
1 2 T 3 T ln CO2 4 T ln i FC
(32)
64
PO2
CO2 (concentration of oxygen at catalytic interface)
498
6
5.08 10 e T
It is difficult to compare the activation loss approximation used in the Lee model to the standard
method since and io are not known for a given set of empirical parameters 1 4 . In
addition, the CO2 parameter in the Lee model represents the concentration of oxygen at the
cathode catalytic layer. To account for cathode GDL properties, only parameter 3 can be
adjusted which may not yield a good representation over all temperatures and air pressures. The
The ohmic loss expression used in the Lee model is essentially the same as the standard
method described by equation 17. The Lee model goes into more detail in how to calculate the
internal resistance of the cell. First, the cell resistance is divided into electronic resistance and
ionic resistance components as shown below in equation 33 (Lee & Wang, 2007).
V ohm ic i R m R c (33)
Second, the ionic resistance of the Nafion membrane (Rm) is expressed in terms of membrane
specific resistivity, m .
m tm
Rm (34)
A
Third, the specific resistivity of the Nafion membrane is given by an empirical expression
(equation 35).
65
i T i FC
2 2.5
Fourth, in the Lee model, there is no value given for Rc (Lee & Wang, 2007). The electronic
resistance of the cell is most likely measured empirically. To compare the ohmic loss between the
Lee model and the standard method is ambiguous. Since the standard method only states a
constant value for the cell resistance, it is only an approximation to the true cell resistance over
Figure 27. MEA Resistance versus Current Density for Different Relative Humidity Values
In figure 27, the cell resistance is plotted against current density for a range of relative
humidity values of the anode gas. Since relative humidity does not affect the electronic resistance
of the cell, it is the ionic cell resistance that is affected. In other words, specific resistance of the
66
Nafion membrane is affected by the stoichiometry and relative humidity of the anode gas.
Although the hydration state of the Nafion membrane depends upon fuel cell design, figure 27
illustrates that the cell resistance is not always constant over a range over current densities. For a
relative humidity of 25%, cell resistance is inversely proportional to current densities under 0.7
A/cm2. In this case, a constant resistance is not a good approximation to cell resistance over the
full range of current densities. On the other hand, at 100% relative humidity, cell resistance is
indeed constant across the full range of current densities. Figure 27 shows a worst case increase
in cell resistance of 60% at 0.3 A/cm2 depending upon relative humidity. The point is that Nafion
membrane specific resistance is affected by stoichiometry and relative humidity of the anode gas.
The Lee model for cell resistance uses the parameter in the specific resistivity expression
(equation 35) to account for stoichiometry and relative humidity of the anode gas. The parameter
can vary between 14 and 23 and is determined by membrane preparation, relative humidity,
and stoichiometry of the anode gas (Lee & Wang, 2007). To visualize the cell resistance as a
function of temperature and current, the following scenario is calculated in Matlab. Assume a cell
51 um. The cell operates over a temperature range of 40oC to 90oC and a range of current
densities from 0.1 A/cm2 to 1 A/cm2. For each integer value of between 14 and23, create a
surface plot. The Matlab code for this cell resistance example is listed in appendix G and the
67
Figure 28. Cell Resistance versus Current Density and Temperature (14 < < 23)
In figure 28, the top surface plot is for = 14 and the surface plots proceed downwards as
increases. The bottom surface plot corresponds to = 23. The surface plots show a cell
resistance dependence upon both current density and temperature (Lee model). The ratio of
highest to lowest cell resistance is about 1.16 to 1. In other words, depending upon the value of
, temperature, and the current density, the cell resistance can increase by 16% in the Lee
model. The Simscape sub-component code for the ohmic resistor is listed in appendix H.
To complete the PEMFC Simscape model, the subcomponents are assembled in a top
level file. However, before the final PEMFC model can be assembled, there are three more sub-
components. First, the concentration loss of the fuel cell is represented in the electrical circuit as
dependent resistor that implements the concentration loss expressed by equation 18. The
Simscape sub-component code for the concentration resistor is listed in appendix I. Second, the
double layer capacitor represents the internal capacitance of the fuel cell due to the
capacitor and the code is listed in appendix J. Third, the flow rate sub-component calculates
68
hydrogen and air flow from fuel cell current. The flow rate sub-component code is listed in
appendix K. The PEMFC model is assembled in the top level file listed in appendix L. The
assembled fuel cell model is then tested in the following Simscape simulation (figure 29).
In the test simulation shown above in figure 29, the PEMFC has a load that increases over time.
The stack voltage and fuel cell current data are extracted which results in the following
polarization curve and power curve plots shown below in figure 30.
The parameters for the PEMFC model are set to mimic the Horizon H-100 fuel cell. On the
polarization curve, two points to check are the open circuit voltage and the nominal operating
point. From the Horizon H-100 fuel cell stack user’s manual, the open circuit voltage should be 19
V and the nominal operating point should be 12 V at 8.3 A (Horizon, 2013). This is in agreement
69
with the PEMFC model polarization curve. On the PEMFC power curve, the maximum power
should be 100 W at the nominal operating point at 8.3 A (Horizon, 2013). In accordance with H-
100 operation, the hydrogen pressure is set to 0.5 atm. and the air pressure is left at 1 atm. The
Note that the number of cells is the same as the Horizon H-100 fuel cell stack (20 cells). Last, the
dynamic behavior of the PEMFC model is tested for a range of double layer capacitance values.
The results for the double layer capacitor ranging from 1 uF to 1 F are shown below in figure 32.
Figure 32. PEMFC Dynamic Response for Range of Double Layer Capacitance
70
3.5 Balance of Plant Model
To properly model the Horizon H-100 fuel cell, the BOP electrical model must be
included. Within the BOP equipment that ships with the fuel cell, only the fuel cell controller, fan,
and temperature sensor consume electrical power. The fuel cell controller is powered by an
external 13 V power supply and it then supplies power to the fan and temperature sensor. The
Input power for the BOP circuit comes from the PMS electrical bus. The DC-to-DC converter then
produces a regulated 13 V to run the fuel cell controller. The variable resistor labeled as
BOP_Load acts as the fuel cell controller electrical load. A control signal adjusts the BOP_Load
as a function of time and the value of the control signal is the BOP_Load resistance.
During operation, the fuel cell controller consumes a somewhat constant amount of power. The
change in BOP_Load over time is due to the fan being on or off. In figure 34, the control signal
simulates the BOP_Load when the fan is normally off and then turns on for a short duration. Note
that the BOP_Load control signal is set to 18 Ω when the fan is off and is set to 11 Ω when the
71
fan is on. This will correspond to a BOP electrical load of roughly 0.76 A when the fan is off and
Most of the electrical load on the power system consists of the ESC. The ESC delivers
power to the BDCM which in turn, spins the propeller. The propeller then converts the rotational
motion of the BDCM shaft into linear thrust. To size the electrical load, the physical properties of
the aircraft and the design goals are required. In this study, the ASU Composite UAV is the
aircraft used for the simulations. One design goal for the power system is to power the aircraft
using only the fuel cell during cruise. It is given that the power required during cruise flight
conditions is 25% of maximum and that the thrust to weight ratio of the aircraft is between 0.2 and
0.4 (Raymer, 1989). Converting the aircraft weight of 5.5 kg to an equivalent gravitational
downward force gives a value of 53.9 Newtons (N) which then leads to the following thrust
The ASU Composite UAV is designed to have a cruising speed of 20 m/s (Krauch et al., 2012).
Using the definition of power (power = force x velocity), the thrust constraint equation can be
expressed in terms of power. The required power during cruising conditions (cruise power) is
Equations 37 and 38 express the amount of thrust power necessary to yield a thrust to weight ratio
between 0.2 and 0.4. To convert from thrust power to the rotational mechanical power at the BDCM
shaft, the propeller efficiency can be used. During this study, different propellers for model aircraft
applications have been examined, however propeller efficiency is not a specification that is reported
in the sources reviewed by this author. Instead, a reasonable estimate for a propeller efficiency
72
63.5 W PMech ,cruise 127 W
(39)
To then convert from mechanical power to electrical power, the BDCM efficiency can be used. At
this point, the BDCM must be selected in order to obtain an accurate efficiency value. In the original
ASU Composite UAV design, three BDCMs were used. All three motors used were from the
manufacturer AXI (Czech Republic); AXI product number 2217/16. In this study, the three BDCM
are replaced by one larger BDCM, the AXI 4120/14. From the AXI application notes, the AXI
4120/14 combined with a 13 x 8 propeller has an efficiency of 84% (“AXI Motor Applications
4120/14”, 2015).
As a result of equation 41, a fuel cell power of 100 W should be sufficient to meet the design goal
of powering the UAV during cruise flight using only the fuel cell. During takeoff and climb portions
of flight, the maximum battery power required will then be in the following range (equation 43).
both Simscape library components and custom Simscape text-based components. The inputs to
the MPSS are power in from the electrical bus and a throttle command that ranges from 0% to
100%. The outputs from the MPSS are the propeller speed setpoint, actual propeller speed,
electrical power consumed, and mechanical power produced at the BDCM shaft. Central to the
operation of the MPSS is the Simscape servomotor component from the SimElectronicsTM add-on
library. The servomotor component represents a BDCM and its controller (the ESC) operating in a
closed loop torque control mode. An outer control loop based on propeller speed is then placed
around the servomotor. A torque load is then applied to the shaft of the servomotor to represent
the aerodynamic load of the propeller. The propeller speed and torque load are adjusted to model
73
the AXI 4120/14 BDCM with a 13 x 8 APC propeller. The MPSS block diagram is shown below in
figure 35.
Using the figures given in the AXI application notes, the AXI 4120/14 motor combined with a 13 x
Table 5
In the MPSS block diagram (figure 35), the throttle input is converted to a power request value
which is then converted to an equivalent propeller speed value. The propeller speed value is then
used as the setpoint to the outer control loop. The outer control loop generates an error signal
from the difference of the setpoint and the servomotor shaft speed and then feeds that error
signal into a PI controller. The output of the PI controller is then converted to a torque value and
fed to the servomotor torque request input. The output of the servomotor is the shaft speed which
is then used as feedback for the outer control loop and also as an input for a load-torque lookup
74
table. The output of the load-torque lookup table is fed to the input of a custom Simscape torque
load component which applies the torque load to the shaft of the servomotor. The Simscape code
for the revolutions per minute (RPM) versus power lookup table, the torque versus RPM lookup
In figure 36, the propeller actual speed matches the setpoint as the throttle ranges from
0% to 100% which demonstrates that the outer control loop is functioning as expected. The curve
of the propeller speed versus throttle graph is determined by the RPM versus power lookup table.
Data for the lookup table was derived from a combination of measured data and the motor
manufacturer’s application note (“AXI Motor Applications 4120/14”, 2015). First, the maximum
point on the propeller speed versus throttle graph is established from the AXI application note,
(100%, 7300 RPM). Second, a test fixture was constructed to measure the propeller speed and
power versus throttle curves of the AXI 1420/14 motor. Images of the thrust test bench (TTB) are
shown below in figure 37. Experimental data measuring ESC waveforms, thrust, motor power,
ESC current, ESC voltage, and propeller RPM were collected during the fall 2015 semester. A
75
Figure 37. Thrust Test Bench
In the TTB setup, the ESC driving the AXI 4120/14 motor was the JETI Advance 70 Pro electronic
speed controller. Two limitations were present during testing. First, the propeller used in the TTB
was different than the 13 x 8 APC propeller modeled in the MPSS block. Instead, the TTB used a
10 x 7 APC prop. Second, the power sources used to drive the ESC/motor combination were
limited in power capacity and could not drive the motor to its full potential. The end result of motor
testing with the TTB was a characterization of the propeller speed versus throttle curve and the
power versus throttle curve over a limited range. Shown below in figure 38 are two graphs of
76
The curve of the propeller speed versus throttle graph in figure 38 is brought over to the propeller
speed versus throttle graph in figure 36 with the curve scaled to fit the maximum point, (100%,
7300 RPM). Similarly, the power versus throttle graph shows a linear relationship which is
brought over to the mechanical power curve of figure 36. The mechanical power curve of figure
36 is determined by the lookup table inside the TorquevsRPM block of the MPSS model. Note
that in the power versus throttle curve of figure 36, the electrical and mechanical curves are
nearly identical for values under 110 W. Similar to the propeller speed versus throttle graph of
figure 36, the maximum electrical and mechanical power points in the power versus throttle graph
are set to the values given in the AXI application note (“AXI Motor Applications 4120/14”, 2015).
The curves shown in figure 36 result from a simulation using the MPSS block while the curves
presented in figure 38 are measured data from the TTB. Although the characteristics between
figures are closely matched, they are not perfectly matched. The interplay between
subcomponents within the MPSS block make a perfect match difficult and therefore MPSS
77
CHAPTER 4
SYSTEM INTEGRATION
In this research, a unique approach to APM system design for UAV applications is
presented. Building upon the concept of the ASU APM circuit, a new APM system design is put
forth that uses a MOSFET approach to control battery current delivered to the electrical bus. The
proposed APM system design has the fuel cell connected directly to the bus while the battery is
connected to the bus through a p-channel enhancement MOSFET. This is similar to the ASU
APM circuit, but in this new circuit design, the MOSFET is used in a linear fashion to control
battery current. Additionally, the fuel cell voltage and current and the battery voltage and current
are monitored. With this information, a supervisory microprocessor can run a digital observer
algorithm to track hydrogen consumption, fuel cell efficiency, battery SOC, and power delivered
from both sources. Due to the circuit features of battery augmented current delivery to the bus
and power source monitoring, the proposed APM circuit is named Current Assist and Observer
Circuit (CAOC). In this APM system design, no DC-to-DC converter is used in the main power
path, thus reducing the APM weight to a minimum. Also, no series diodes are used at the power
sources as in the ASU APM circuit, so no energy is lost to the forward voltage drop of the series
diodes. During the cruise phase of flight, in steady-state operation, the power loss of the CAOC
drops to near zero yielding a better power efficiency than some DC-to-DC converter approaches
(compared to figure 14). Figure 39 shown below displays a conceptual diagram of the CAOC.
78
Figure 39. CAOC Concept
Resistors R1 and R2 are current sense resistors with extremely low resistance values of 5 mΩ.
Due their extremely low resistance, the voltage drop and power dissipated by R1 and R2 is small,
almost negligible. Operational amplifiers (OpAmps) U1 and U2 measure the current from the fuel
cell and battery respectively. They take the differential voltage drop across R1 and R2 and
amplify the voltage (e.g. voltage gain factor around 20). The OpAmps convert the measured
differential voltage to a single-ended voltage which can then be sent to the observer. The output
voltage of U1 is also sent to the inverting input of U3. OpAmp U3 takes the difference between a
setpoint voltage and the output of U1, amplifies the difference, and controls the gate voltage of
Q1. MOSFET Q1, normally in the off-state due to resistor R3, begins to conduct only when the
gate voltage is driven lower than the source voltage minus the MOSFET threshold voltage
(Vthreshold). The setpoint voltage is generated by the supervisory microprocessor and is a voltage
command for the fuel cell current setpoint. When the actual fuel cell current is below the setpoint,
Q1 remains in the off-state. When the fuel cell current increases past the setpoint value, U3
drives the gate voltage of Q1 to a lower voltage, thus forcing Q1 to conduct slightly. The amount
that Q1 conducts is then proportional to the difference between the actual fuel cell current and the
setpoint value. The end result is that the fuel cell current is held at a constant value fixed by the
79
setpoint voltage. If the bus current goes above the fuel cell current setpoint, extra current is
supplied by the battery. If the bus current is below the setpoint, then the fuel cell current will
decrease to the bus current and the battery remains disconnected from the bus. Last, the fuel cell
voltage and battery voltage are monitored and sent to the observer. Note that the MOSFET Q1
has to handle the maximum current delivered by the battery. Device selection for Q1 must have a
continuous current specification larger than the maximum battery current plus a 50% safety
factor. In addition, special consideration must be given to the heat sinking of Q1 so that the
junction temperature stays below the maximum allowed value. A side benefit of the CAOC
method is that the setpoint may be lower than the bus current during cruise conditions. In this
case, the battery will contribute part of the cruise mode current. In this manner, the battery can
contribute to maximum flight endurance and is a new option for mission planners.
the CAOC is combined with a microprocessor. The circuit components used for the CAOC are
pulled from the SimElectronics library while the microprocessor is represented by a Stateflow
block. Stateflow is a separate Matlab product that functions inside the Simulink environment to
create flow charts and state machines. The Stateflow block acts as the supervisory
microprocessor by processing output data from the CAOC and controlling the setpoint voltage
80
Figure 40. HPMC Circuit
In figure 40, the tan colored block labeled PM_Controller is the Stateflow block that models the
supervisory microprocessor. Inside the Stateflow block, the APM control algorithm can be
implemented as a state machine or flow chart. From this point forward, the Stateflow block will be
referred to as the microprocessor block (MB). Components R1, R2, R3, U1, U2, U3, and Q1
make up the CAOC and perform the same function in figure 40 as they do in figure 39. Note that
the inputs to the MB are the fuel cell and battery currents and their integrated values. The MB
does not monitor the fuel cell voltage or the battery voltage; only the current values are necessary
for power management control. During flight, the throttle sets the motor power level as a
percentage. During takeoff and the initial climb, the throttle is set to 100% and full power
delivered to the motor. When the desired cruising altitude is reached, the aircraft levels out and
the throttle is reduced to 25%. Note that the MB does not use the throttle as an input. In this
scheme, power management does not require information about the throttle setting.
Inside the MB, there are three flow charts that define a simple power management
scheme used during flight. The first two flow charts act as coarse fuel gages for the battery SOC
and the hydrogen fuel level. They divide the battery SOC and hydrogen fuel level into three
81
Figure 41. Battery and Hydrogen Coarse Fuel Gages
For the two fuel gage flow charts shown in figure 41, the integrated battery and fuel cell currents
track the amount of electrical charge delivered from the two sources. For the battery state, charge
delivered from the battery is divided by 3600 before it is fed into the MB which changes the units
from coulombs to amp-hours. The amp-hours delivered from the battery is divided by the battery
capacity and then the ratio is subtracted from the initial SOC. When the battery energy reserve
drops below 90% of battery capacity, the battery state changes from full to mid. Likewise, when
the battery energy reserve drops below 40%, the battery state changes from mid to low. Note that
the battery state can also increase from low to mid to full. Since the battery can be recharged
during flight, the battery state must be able to change in both directions. For the hydrogen fuel
state, the charge delivered from the fuel cell is fed into the MB. The corresponding amount
n q
Liters 24.46 (44)
2F
In equation 44, n is the number of cells in the fuel cell stack, q is the amount of charge delivered
from the fuel cell in units of coulombs, and F is Faraday’s constant. The factor 24.46 is the
number of liters per mole for any ideal gas at STP. The amount of hydrogen consumed is
subtracted from the total amount of hydrogen stored. When the amount of hydrogen fuel drops
below 80% of the total hydrogen initially stored, then the fuel state drops from full to mid. When
the hydrogen fuel level drops below 30%, the fuel state drops from mid to low. Note that the fuel
82
state can only decrease since the aircraft cannot refuel during flight. The third flow chart controls
the setpoint voltage to the CAOC. The setpoint is a voltage command to the CAOC which
controls the current delivered from the fuel cell. If the load demands more current than the fuel
cell setpoint level, then the battery automatically supplies the extra current required. When the
setpoint is at a maximum value of 9 V, then the fuel cell delivers current up to its nominal
operating point value before the battery delivers any current. For this model, the fuel cell
parameters are set to model the Horizon H-100 fuel cell stack and the nominal operating point is
8.3 A at 12 V. In other words, with a setpoint of 9 V, the fuel cell will supply load currents up to 8.3
A without any help from the battery. When the load requires current levels above 8.3 A, the
battery will augment the fuel cell current. As the setpoint is lowered, the battery will begin to
deliver current at lower load currents. For example, with the setpoint at 7 V, the fuel cell will
deliver up to 6.3 A; the battery will supply any load current above 6.3 A. Shown below in figure 42
In the flow chart displayed in figure 42, the setpoint starts out at a maximum value of 9 V.
At the beginning of a flight profile, maximum power is needed from both fuel cell and battery to
satisfy the power demands of 100% throttle. The exit condition from the leftmost block labeled
Normal is when the battery current drops below 0.1 A. This keeps the setpoint at 9 V during the
takeoff and climb phases of the flight profile. When the cruising altitude is reached, the throttle is
reduced from 100% to 25% at which point the fuel cell can supply the total load current and thus
the battery current drops to zero. Transition from the Normal block to the Cruise block then lowers
the setpoint to a level of 7.5 - h. The variable h is set by a parameter called hybrid which can vary
between 0 and 2. The parameter hybrid controls the ratio of fuel cell power to battery power
during cruise flight. In this manner, the battery can augment fuel cell power during cruising
83
conditions and extend flight endurance by trading battery power for hydrogen fuel. If during cruise
flight, the battery enters the low state, then the setpoint is raised to 7.5 by leaving the Cruise
block and entering the FConly block. This sets the fuel cell current level to the expected load
during cruise flight. Note that the FConly block works in conjunction with the Delta_1 block to
raise the setpoint to 7.5 V at a controlled rate. Once the setpoint reaches 7.5 V, the battery
current will drop to zero and the battery SOC will decrease no further under normal cruise load.
The state machine will also leave the cruise block when the cruise phase of the flight profile is
complete. At this point, the battery still has enough energy reserve to assist the fuel cell with
temporary loads greater than the expected cruise load for events such as wind gusts, aerial
maneuvers, or taxiing.
To model the proposed UAV hybrid power system, the Simscape components are
assembled in a top level file. The fuel cell and the battery are connected to the electrical bus
through the HPMC. Two electrical loads, the motor-propeller model and the BOP model, are
connected directly to the bus. The battery charger receives power from the bus and its output is
connected to the battery. In addition, two observer blocks are used to monitor voltages, currents,
power, energy consumed, system state, and fuel cell model internal values. The diagrams for the
Observer_1 and Observer_2 blocks are listed in appendices O and P respectively. The model
configuration parameters choose the ode15s solver with a reduced relative tolerance of 10-6. A
screen capture of the model configuration parameters is displayed in appendix Q. The top level
84
Figure 43. Hybrid Power System Model
The throttle input to the hybrid power system is realized using a Simulink signal builder block at
the top level named Flight_Profile. The signal stored is the flight profile; that is, the throttle
settings during the takeoff, climb, cruise, and landing phases of flight. A second signal builder
input to the model is contained within the Balance_of_Plant block named the Fuel_Cell_BOP.
This signal represents the BOP load to the electric bus and is expressed directly in units of ohms.
The final input to the model is a signal builder block contained within the Controller named
BAS_signal. It activates the current assist feature of the HPMC. When the BAS_signal is zero,
current assist is not active. In this case, the fuel cell will supply power up to its nominal operating
point before the battery assists with excess load current demands. When the BAS_signal is one,
current assist is active and the hybrid parameter may lower the current assist threshold, allowing
the battery to augment the fuel cell at a lower power point than the nominal.
In 2012, the ASU Composite UAV was originally designed for electric powered flight
using a 5 Ah LiPo-3S battery pack (Krauch et al., 2012). The following year, a hybrid power plant
study investigated the use of a hybrid fuel cell and battery power system to potentially increase
the ASU Composite UAV flight endurance from 25 minutes (battery only), to 6 hours (Monaco,
2013). In this study, a hybrid power system for the ASU Composite UAV is now modeled in
85
Simscape. The power system simulation demonstrates both the functionality of the HPMC and
determines the amount of hydrogen fuel required to realize a 6 hour flight time.
In this section, the flight time is set to 6 hours with a typical flight profile and BOP load.
The flight profile starts out with a 3 minute interval at 100% power to simulate the UAV starting on
the runway and taking off. The next 4 minutes are set to a 60% power level which represents the
initial climb to cruising altitude. The UAV then cruises for 58 minutes at 25% power. At 65 minutes
into the flight, the UAV climbs again for 4 minutes at a shallow climb rate requiring a power level
of 40%. The UAV then cruises again for 45 minutes at 25% power. Wind gusts appear at the 114
minute time mark and persist for 7 minutes. During the wind gusts, the throttle alternates between
25% and 35% power to simulate brief intervals of increased power for course corrections. The
UAV then finds clear skies and smooth flight conditions to cruise for another 3.8 hours at 25%
power. At the 5.9 hour mark, the UAV is near the landing strip and temporarily increases power to
30% to line up for final descent to the runway. Power then decreases to 20% for 4 minutes during
the final descent and landing phases of the flight profile. Once on the ground, the total powered
flight time has been 357 minutes; almost 6 hours. The UAV then increases power to 30% to taxi
off the runway and return to its designated maintenance area. Finally, the UAV idles at 10%
power for a minute before shutting down. In total, the flight profile simulates 6 hours of power
delivered to the electric motor from the hybrid fuel cell/battery power system.
86
There are three input signals to the simulation as shown above in figure 44. The first
signal is the throttle percentage which defines the flight profile. The second input signal is the
BOP load. The actual BOP load signal is defined in terms of ohms (section 3.5), but is estimated
in Watts in figure 44 for a hybrid value of zero. The fuel cell BOP signal sets the nominal BOP
load near 10.4 W. At four points during the flight, a 3 minute interval of increased BOP load to
roughly 16.9 W occurs to simulate the fuel cell fan activating. The third input signal is the BAS
signal which is either off (value = 0) or on (value = 1). In figure 44, the BAS signal is multiplied by
8 to make it easier to see the logic level on the graph. With the BAS_signal, current assist is
activated just before the 9 minute mark, maintained in the on state during the flight, and then
In this first group of two simulations, the hybrid parameter is first set to zero and then to a
value of 1.5. When the hybrid parameter is zero, battery usage is minimized and hydrogen fuel
consumption is maximized. The fuel cell current threshold for battery assist is 6.8 A. In other
words, for electrical loads less than 6.8 A, the fuel cell will supply the total current delivered to the
load. If the electrical load goes above 6.8 A, then the battery contributes the extra load current.
When the hybrid parameter is set to a value of 1.5, the fuel cell current threshold for battery assist
is lowered to 5.3 A. That is, the fuel cell will deliver load currents up to 5.3 A while the battery will
deliver load currents above 5.3 A. The results of the two simulations are plotted in figures 45
through 47 for comparison. Measured runtime for each simulation is 25 seconds. The simulations
show that 366.5 liters of hydrogen are consumed when the hybrid parameter is set to zero and
327 liters are consumed when the hybrid parameter is 1.5. The battery final SOC is 72.1% when
the hybrid parameter is zero and 35.2% when the hybrid parameter is 1.5. Adjusting the hybrid
parameter from zero to 1.5 has effectively traded 36.9% battery SOC for 39.5 liters of hydrogen.
In the graph titles of figure 46, total energy is listed as 583.5 Wh for the hybrid parameter set to
zero, and 585.2 Wh for the hybrid parameter equal to 1.5. These energy values are the total
energy delivered from the fuel cell and battery during the flight profile. Note that there is a 1.7 Wh
increase in energy delivered from the sources when the hybrid parameter is increased to 1.5. The
87
energy increase is expected to be attributed to extra power dissipated by the HPMC transistor
Q1.
Figure 45. Fuel Cell and Battery Voltage and Current versus Time (hybrid = 0 and 1.5)
88
Figure 47. Battery SOC and Hydrogen Consumption versus Time
To automate the setting of model parameters and input signals, a Matlab script is used.
Within the flight profile script, the hydrogen pressure is set to 0.5 atm and the air pressure is set
to 1 atm. Other parameters set by the script are the total hydrogen fuel stored onboard, battery
capacity, battery initial SOC, number of cells in the fuel cell stack, and the hybrid parameter. Input
signals to the simulation are set using the Matlab signal builder function. The flight profile script is
listed in appendix R. Matlab scripts to generate figures 45, 46, and 47 are listed in appendices S,
T, and U respectively.
Inside the HPMC, transistor Q1 acts as a series regulator to control the battery current
delivered to the bus. When the battery current is zero, no power is dissipated by Q1; however, as
the battery current increases, so does the power dissipated by Q1. Shown below in figure 48 is a
closer look at the power consumed by transistor Q1. The Matlab code to generate figure 48 is
listed in appendix V.
89
Figure 48. Q1 Power Dissipation versus Time
From figure 48, the peak power dissipated by Q1 is 107 W which occurs at the 140 second time
mark. From the flight profile, the throttle is held at 100% from 140 seconds to 320 seconds. Q1
power decays during this time interval while the propeller spins at its top rotational speed of 7300
RPM. As the throttle is reduce from 100%, the battery current decreases and so does Q1 power.
With the hybrid parameter at zero, Q1 power drops to zero when no current assist is needed from
the battery. In this case, the HPMC draws minimal power; only the quiescent power necessary for
the electronics. When the hybrid parameter is 1.5, Q1 cruise power is 1.29 W and the battery
delivers a cruise assist current of 956 mA. During this time, the voltage dropped across Q1
(source relative to drain) is Vsd = 1.35 V and the effective source-to-drain resistance is 1.41 Ω.
If the power supplied by the sources, the power dissipated by Q1 and the load are
integrated over the entire flight profile, the total energy of the sources and loads can be found.
Taking the difference of the total energies between the hybrid = 1.5 and hybrid = 0 simulation
runs provides further insight into the operation of the system. Subtracting the total energies of the
hybrid = 0 run from the hybrid = 1.5 run gives the following results: the sources produce 1.7462
Wh more energy, Q1 dissipates 3.4157 Wh more energy, and the motor-propeller load consumes
1.4339 Wh less energy. The expected result is that the change in Q1 load energy plus the
change in motor-propeller load energy should equal the change in source energy. Note that the
90
motor-propeller load consumes less power, so the change in motor-propeller load energy is
3600 J
1.9818 Wh 1.7462 Wh 848.16 J (46)
Wh
Equation 46 shows that there is an 848 Joule discrepancy in the change of total energy between
simulation runs. When compared to the total energy consumed during the hybrid = 1.5 simulation
run, this amounts to a 0.04% error. The Matlab script which calculates the change in total
energies between runs is listed in appendix W. Table 6 shown below lists the total energies of
Table 6
In table 6, the loads consist of the motor-propeller load, Q1 load, and BOP load; the sources are
the fuel cell and the battery. Note that for the two simulations, the sources produce about 3 Wh of
extra energy that is not accounted for in the listed loads. The extra energy is consumed by the
two DC-to-DC converters in the BOP block and the battery charger block. Note in figure 45, the
bus voltage is 14.50 V when the current assist is active and the bus voltage drops to 13.65 V
when there is no current assist. During the h = 1.5 simulation run, the bus voltage is at the higher
voltage level of 14.5 V for about 3 hours. Note that the excess energy for the h = 1.5 run is lower
than that of the h = 0 run and the difference between the two values accounts for the 848 J
discrepancy. Also, having calculated the total energy dissipated by Q1 and the total energy
produced by the sources, power system efficiency can be calculated as the ratio. For the h = 0
91
and h = 1.5 runs, power system efficiency excluding the BOP and battery charger loads is
To reach a flight time of 6 hours for the ASU Composite UAV, the two previous
simulations calculated the amount of hydrogen required to be 366.5 L without current assist and
327 L with current assist. If the hydrogen storage tank onboard the UAV is the Ninja Pro v2
90/4500 filled to a pressure of 248 bar as mentioned in section 2.9, then the total amount of
hydrogen available is only 298.9 L. The LiPo battery used during the first two simulations was
modeled after the MaxAmps LiPo 4S 8Ah battery. Two possibilities to increase the flight time to 6
hours are to increase the hydrogen pressure of the storage tank or to use a larger battery. For the
first option of increasing the hydrogen storage tank pressure, the 20% safety factor mentioned in
section 2.9 can be decreased. The hydrogen storage tank is approved for air pressures up to 310
bar. The problem is that the storage tank is designed to hold compressed air and not hydrogen.
Decreasing the safety factor from 20% down to 12% may not be safe. However, if the storage
tank is filled to a pressure of 272.8 bar, then 328.8 L of hydrogen can be stored. Using the current
assist feature and filling the storage tank to 272.8 bar would result in a 6 hour flight time. For the
second option of a larger battery, a second group of simulations varies the LiPo battery capacity
from 8 Ah to 12 Ah and determines the amount of hydrogen consumed during each simulation.
Results for this group of five simulations are shown below in figure 49 and the Matlab code is
listed in appendix Y. The amount of hydrogen fuel consumed during each simulation run is listed
in the graph legends. For example, a line in the graph legend lists the text “Battery: 8Ah,
H2:327.2”. This translates to 327.2 liters of hydrogen consumed over the six hour flight profile for
the case where the battery capacity equals 8 Ah. The hybrid parameter is set to 1.5 for all
92
Figure 49. Battery SOC and Hydrogen Consumed for Different Battery Capacities
From figure 49, increasing the battery capacity to 12 Ah is necessary to keep the hydrogen
consumption under 298.9 L. Table 7 listed below shows the total hydrogen consumed and final
battery SOC for each of the battery capacities simulated over the six hour flight profile.
Table 7
Hydrogen Consumed and Battery Final SOC for Different Battery Capacities
Battery Capacity Hydrogen Consumed Hydrogen Consumed Battery Final SOC
(Ah) (L) (g) (%)
8 327.2 27.0 35.2
9 319.0 26.3 35.4
10 311.6 25.7 36.3
11 304.1 25.1 36.9
12 296.8 24.5 37.5
Table 8
93
From table 8, changing from the LiPo 4S 8 Ah battery to the 12 Ah battery increases the power
system weight by 346 g. Also note that the 11 Ah battery has a significantly lower battery weight
of only 850 g. If the 11 Ah battery could be used, then the power system weight would only
increase by 86 g. The weight savings of the 11 Ah battery over the 12 Ah battery is probably due
to the lower discharge rating. From figure 45, the maximum battery current is under 25 A; well
below the 40 C discharge rating of the 11 Ah battery. If the hydrogen tank pressure is increased
to 254 bar (18% safety factor), then 306 L of hydrogen is available and the 11 Ah battery is
sufficient to complete the six hour flight. Due to the reduced weight of the 11 Ah battery, and only
a 2% compromise to the hydrogen storage tank safety factor, this third option is the most
attractive. Figure 50 shown below displays the power system performance during the six hour
In figure 50, the controller deactivates the current assist at the 300 minute time mark due to the
battery SOC entering the Battery_Low state (40% SOC). At this point, the CAOC setpoint is
increased from 6 to 7.5 at a controlled rate (see figure 42). Note that the bus voltage (fuel cell
voltage) decreases from 14.5 V to 13.65 V during this transition. Since the fuel cell voltage is
higher when the current assist is active, an added benefit of increased fuel cell efficiency is
realized. Using the higher heating value (HHV) of hydrogen, the fuel cell efficiency is 49% with
current assist and 46% without. If referencing the lower heating value (LHV) of hydrogen, then the
fuel cell efficiency is 58% with current assist and 54.6% without. To examine the current and
94
power graphs from figure 50 more closely, figure 51 shows the same data with the y-axis ranges
changed.
Figure 51. Hybrid Power System Performance for 11 Ah Battery Expanded View
To summarize, system level simulations have shown that with 306 L of hydrogen and an 11 Ah
LiPo battery, a six hour flight of the ASU Composite UAV is possible using the current assist
feature of the HPMC. Without current assist, 366.5 L of hydrogen would be necessary to
complete the six hour flight. Using battery power during cruise conditions has in effect, reduced
the amount of hydrogen required by 60.5 L and has transferred about 74 Wh of energy load to the
battery. The battery SOC at the end of the flight is 79.7% without current assist and 36.9% with
current assist. The difference between these battery SOC values is due to the battery delivering
the extra 74 Wh of energy required to complete the six hour flight profile.
When excess hydrogen is available, the battery charger circuit can be used to recharge
the battery in flight. Mission planners may want to recharge the battery in flight to reach higher
altitudes or to reduce maintenance time on the ground. Although the battery charger is already
integrated into the hybrid power system model (see figure 43), the controller SP_Block (see figure
42) must be altered. With the battery charger, the controller SP_Block is replaced by the
95
Figure 52. CTRL_Block Flow Chart
The Stateflow default transition is the starting point for the CTRL_Block flow chart and enters the
Normal block on the left side. The CAOC setpoint voltage is assigned the value stored in the
flowchart variable Setpoint. Starting at the Normal block, a working variable named csp (current
setpoint) is initialized to a value of 9 which is then assigned to Setpoint. When the variable
Setpoint is updated, the CAOC setpoint voltage is updated to a physical value of 9 V. When the
battery current drops below 0.1 A and the BAS signal has a value of one, then the state machine
transitions from the Normal block to the Ctrl_Mode block. The Ctrl_Mode block simply determines
if the battery charger is active during the simulation run. If the battery charger is active, then
variable ca (charger armed) is one, otherwise it is zero. Note that the flow chart logic of the
SP_Block is contained within the CTRL_Block. At the Ctrl_Mode block, the transition branch
where ca = 0 goes to the same flow chart as the SP_Block. In other words, when the battery
charger is not active (ca = 0), the same SP_Block flow chart logic is used. The value of ca is set
in the Matlab environment by the flight profile script. If the value of ca is set to one, then the
battery charger is active and the state machine will transition from the Ctrl_Mode block to the
Charger_Gate block. The Charger_Gate block then waits for the BAS signal to equal a value of
two before it activates the battery charger. Here, the BAS signal is used for two purposes. When
the BAS signal is zero, neither current assist or battery charger are active. With a BAS signal
value of one, the current assist is active but the battery charger is not. When the BAS signal has
a value of two, both current assist and battery charger are active. The BAS signal controls at
which time intervals during the flight current assist and the battery charger are active. Shown
96
below in figure 53 is a graph of the input signals to the hybrid power system model for the six
hour flight profile. When the BAS signal changes value from one to two, the state machine
transitions from the Charger_Gate block to the Charger_On block. Within the Charger_On block,
the BatCharge variable controls the rate of battery charging and is set to variable cv which is
assigned in the Matlab environment by the flight profile script. A BatCharge value of 0.8
corresponds to a battery charge current of 302 mA. The battery charger will remain active until an
appropriate exit condition is reached. Exit conditions include the BAS signal changing value to
one or zero, or the fuel cell current exceeding 8 A, or the battery charge state entering the
Battery_Full state, or the hydrogen fuel state entering the Fuel_Low state. Upon any valid exit
condition, the state machine transitions from the Charger_On block to the Charger_Off block.
Near the end of the flight profile, the BAS signal will change to zero and the state machine will
transition from the Charger_Off block to the FConly block. The CAOC setpoint is then raised to
In figure 53, the BAS signal value is multiplied by 4 to make it easier to see the logic level on the
graph.
Transitions between the Charger_On block and the Charger_Off block are two way.
Assuming that the fuel state is above the Fuel_Low state and battery charge state is below the
Battery_Full state, the fuel cell current must be under 8 A for the battery charger to remain active.
If the fuel cell current goes above 8 A, then the battery charger is deactivated and may reactivate
when the fuel cell current drops below 8 A. This limit on the fuel cell current prevents overloading
97
of the fuel cell and guarantees the bus voltage will remain above 12 V. If this safeguard was not
in place, then current assist events that occur during battery recharge may overload the fuel cell
which would cause the bus voltage to drop below its minimum operating voltage. A second
condition that may cause the state machine to toggle between the Charger_On block and the
Charger_Off block is battery SOC. To place an upper limit on the battery SOC, the Battery_Full
state signals that the battery charge is complete. If during flight, the battery charge state drops
below 90% due to current assist or throttle commands, then the charger may reactivate and
Figure 54. Fuel Cell and Battery Voltage and Current during Inflight Battery Recharge
To evaluate possible flight times for various hybrid power system configurations, the six
hour flight profile can be used. The flight profile is partitioned into three sections, takeoff, cruise,
and landing. During cruising conditions, the power consumed can be considered constant if wind
gusts and BOP variations are ignored. During takeoff and landing portions of the flight profile,
power is not constant, but the total energy consumed is. Integrating load power during takeoff and
98
landing sections of the flight profile determines the total energy consumed. Power system
configurations can then be evaluated in terms of total energy required. The total energy used
during takeoff and landing is considered constant while the power used during cruise is in terms
of energy per unit time. The total energy a hybrid power system configuration can deliver (EHPS) is
then used to determine the maximum flight time possible. The takeoff and landing energy (ETOL)
is subtracted from EHPS and the amount of energy left over determines how long the UAV can
remain at cruise conditions. First, the electrical power consumed from the motor-propeller block is
added to the power used by the BOP block. The total electrical load energy is then plotted in units
of Watt-hours per minute (Wh/min). Figure 56 below displays the power consumed during the six
From figure 56, the cruise power is found to be 1.51 Wh/min. From the flight profile, the takeoff
segment occurs from time zero to 480 seconds and the landing segment occurs between 21200
seconds and 21800 seconds. Integrating the power during takeoff and landing gives the total
takeoff energy as 31.9 Wh and the total landing energy equal to 13 Wh (ETOL = 44.9 Wh). The
Matlab script that calculates these energies can be found in appendix Y. Flight time is then
calculated in terms of total energy available from the hybrid power system configuration.
99
t [ EHPS 44.9 / 1.51] 18 (47)
Note that takeoff accounts for 8 minutes and landing takes 10 minutes. In equation 47, since the
energy for takeoff and landing is subtracted from the energy available (EHPS – ETOL), the 18
minutes for takeoff and landing must be added to the total flight time. To calculate EHPS, the fuel
cell and battery energies are added together. The amount of energy the fuel cell produces is
equal to the amount of hydrogen energy available times the fuel cell efficiency. For one liter of
To approximate the fuel cell efficiency during the flight, the average between the two efficiencies
obtained with current assist and without is 56.3% (LHV). Therefore, to convert from liters of
hydrogen available (LHA) to total watt hours delivered by the fuel cell (EFC), equation 49 shown
2.7528 Wh
EFC LHA 0.563 1.55 LHA (49)
L
For the total energy delivered from the battery (EBAT), the battery capacity (CBAT) in Ah is
multiplied by the battery voltage. From figure 50, the average battery voltage during flight is
approximately 15.7 V. Assuming that the battery starts out at 100% SOC and ends at 36% SOC
(see table 7), there is 64% of the battery capacity available. Then, to convert from battery
capacity to total energy available from the battery, equation 50 shown below can be used as an
approximation.
100
Substituting equation 51 into equation 47, the flight time is calculated as a function of hydrogen
To check the validity of equation 52, simulation data from figure 49 can be used. Actual flight time
for the six hour flight profile is 21800 seconds or 363.3 minutes. Figure 49 lists the simulation
results for the amount of hydrogen consumed over a range of LiPo 4S battery capacities. Using
this data, the calculated flight time using equation 52 is compared against the actual flight time.
Table 9
From table 9, the error between calculated flight time and actual flight time is under 4%. Note that
the calculated flight time is larger than the actual flight time of 363.3 minutes. Cruise power is
considered constant and variations in cruise power due to wind gusts and BOP load changes are
ignored. The approximation of equation 52 assumes a fuel cell operating at 56.3% efficiency, a
LiPo 4S battery, current assist is active during cruise, and 64% of battery capacity consumed. For
the case where no fuel cell is present and only a battery is used, then LHA can be set to zero.
To evaluate hybrid power system performance for different configurations, power system
weight and flight time are two useful metrics. Using only a battery, the power system weight is
lower than a hybrid configuration, but flight endurance is limited. Using only a 100 W fuel cell will
not work for the ASU Composite UAV since the takeoff power required is 400 W (see figures 46
and 48). Instead, if a 500 W fuel cell were used, then only the fuel cell is required onboard. The
500 W fuel cell startup could be powered externally before takeoff. Using commercially available
LiPo batteries and compressed air tanks, a comparison of weight versus flight time is graphed for
101
combinations of power system components. Total power system weight includes contributions
from the HPMC, fuel cell stack, fuel cell controller, hydrogen storage tank, pressure regulator, and
Table 10
Three commercially available fuel cells are from Horizon Fuel Cell Technologies
(www.horizonfuelcell.com). The H-100 and H-500 fuel cells are for general use and are not
weight optimized. The AST01-01 is a 200 W fuel cell that is optimized for low weight aviation
applications.
Table 11
State of the art carbon fiber compressed air tanks are available from www.ansgear.com. All three
Ninja Pro v2 tanks are rated for air pressures up to 310 bar. For hydrogen storage, a 20%
reduction in maximum pressure is assumed as a safety factor. The listed hydrogen storage
capacities in table 12 shown below are calculated at a storage pressure of 248 bar.
102
Table 12
On all Ninja Pro v2 storage tanks is a high pressure regulator which steps down the tank pressure
from 248 bar down to 27.6 bar. The weight of this first stage regulator is included in the tank
weights listed in table 12. A second stage pressure regulator is necessary to further reduce
hydrogen pressure down to 0.52 bar required by the fuel cell. A commercially available pressure
regulator is the Beswick miniature pressure regulator which has a weight of 0.0062 kg (“Beswick
Pressure Regulator”, 2016). The weight of the HPMC is assumed to be 0.032 kg.
There are 9 power system configurations considered, two conventional (non-hybrid) and
seven hybrid. The two non-hybrid configurations are battery only and fuel cell only. The first
configuration is with the battery only and three battery capacities are chosen: 6 Ah, 11 Ah, and 17
Ah. Equation 52 is used to calculate the expected flight time for each battery capacity (LHA = 0).
A line representing this power system configuration is then constructed on a weight versus flight
time graph as follows. For the lowest battery capacity, the weight is the smallest of the three
batteries. This is the y-coordinate of the first point. Minimum flight time is assumed to be 18
minutes since this is the amount of time for takeoff plus landing. This is the x-coordinate of the
first point. For the second point, the calculated flight time of the 6Ah battery is used as the x-
coordinate while the y-coordinate remains the same. For the third point, the next larger weight of
the 11 Ah battery is used as the y-coordinate while the x-coordinate remains the same as the
second point. The fourth point uses the calculated flight time of the 11 Ah battery as the x-
coordinate while the weight or y-coordinate is the same as the third point. The fifth point keeps
the x-coordinate of the fourth point and uses the 17 Ah battery weight as the y-coordinate. Last,
the sixth point uses the calculated flight time of the 17 Ah battery as the x-coordinate and the y-
coordinate is the same as the fifth point. The end result is a stair-step line with three steps on the
103
graph. Each step represents the minimum weight for the power system configuration for a range
of possible flight times. Next, consider the other non-hybrid configuration which uses the H-500
fuel cell without a battery. The construction of this line is similar to the battery only line except the
weight increase is not due to larger battery sizes; instead the weight increase between steps is
due to larger hydrogen storage tank sizes. The three hydrogen storage tanks considered are the
Ninja Pro 50, 68, and 90. The resulting H-500 only line is again a stair-step line with three steps.
Next, the seven hybrid power system configuration lines are constructed in similar manner to the
H-500 line; but in these cases both fuel cell and battery are present. Six of the hybrid
configurations are combinations of the H-100 or AST01-01 fuel cells combined with the 6 Ah, 11
Ah, or 17 Ah batteries. In all six of these hybrid configuration cases, the current assist feature of
the HPMC is active. Last, the seventh hybrid configuration is labeled the conventional hybrid
configuration where a DC-to-DC converter is present and no current assist is used. For this
conventional hybrid case, a minimum battery size necessary for takeoff and landing is chosen;
the 6 Ah battery. The weight of a 400 W DC-to-DC converter (0.22 kg) is added to the power
system weight which represents a conventional approach for hybrid power systems management.
This conventional hybrid case is special in that it provides a baseline of comparison for the
transistor-based, current assist power management configurations. Since the conventional hybrid
configuration does not use current assist and it includes the weight of a DC-to-DC converter, the
benefits of the transistor-based HPMC using current assist can be noted by comparison. The end
result is a graph of the power system weight versus flight time as shown below in figure 57. The
Matlab script to calculate and plot flight times in figure 57 is listed in appendix AA. The flight time
calculations using equation 52 are approximations only. They assume that the aerodynamic
characteristics of the ASU Composite UAV do not change when the weight changes. In actuality,
the energy used during takeoff and landing as well as cruise power will vary as weight changes.
Additionally, the distribution of weight within the ASU Composite UAV fuselage will affect the
aerodynamic performance of the airframe. Therefore, the flight time data presented in figure 57
should be considered as a rough estimate only, not accounting for aerodynamic performance
changes. All flight time data in figure 57 is applicable to the ASU Composite UAV only.
104
Figure 57. Flight Time versus Weight for Power System Configurations
In figure 57, there are three thin black lines that extend diagonally across the graph.
These are the battery extrapolation lines. These lines represent the weights and flight times
expected when using multiple batteries. The left-most battery extrapolation line represents the
case where multiple 6 Ah batteries are used without a fuel cell. Note that there are 8 circle
markers on this line; each marker represents an additional 6 Ah battery. A total of eight 6 Ah
batteries (0.57 kg each) can fit on the graph where the upper weight limit is 5 kg. Similarly, the
right-most battery extrapolation line represents the case where multiple 17 Ah batteries are used
without a fuel cell. There are 3 circle markers on this line which are the number of 17 Ah batteries
(1.294 kg each) that are under 5 kg. Between the left and right battery extrapolation lines is a
sector on the graph where battery only performance can be expected (battery only sector).
The six non-conventional hybrid configurations are cases that use the transistor-based
HPMC presented in this study. In all six cases, the current assist feature of the HPMC is used.
This group can be divided into two sub-groups; one sub-group using the H-100 fuel cell (solid
lines) and the second using the AST01-01 fuel cell (dashed lines). Each sub-group uses the
same color code for the battery capacity present in the configuration (blue = 6 Ah, red = 11 Ah,
105
green = 17 Ah). The two sub-groups have the same flight times, but the AST01-01 sub-group is
shifted downwards due to the lower weight of the AST01-01 fuel cell.
From figure 57, a comparison between the conventional hybrid configuration (bold purple
line) and the H-100 using the 6 Ah battery hybrid configuration (blue line) can be made. Both
power system configurations have the same fuel cell and battery; only the power management
controllers are different. The conventional hybrid system uses a DC-to-DC converter to control
the power split between fuel cell and battery. The non-conventional configuration (blue line) uses
the transistor-based HPMC approach discussed in sections 4.1 and 4.2. A weight savings of 0.22
kg shifts the blue line down from the purple line on the graph. Due to the current assist feature of
the HPMC, the blue line is also shifted to the right by 40 minutes compared to the purple line. The
gain in flight time is due to the remaining battery energy available after takeoff and landing
energies are subtracted out. In other words, the conventional hybrid configuration does not have
current assist and cannot take advantage of excess battery energy available.
The graph presented in figure 57 can be divided into three regions. First, the region on
the graph between the battery extrapolation lines is labeled the battery only sector. The region to
the left of the battery only sector is where power system performance is worse than using battery
power only. In this context, the term power system performance can be defined as the ratio of
flight time over power system weight (minutes per kg). This first region to the left of the battery
only sector will be labeled region one. Within the battery only sector, power system performance
is roughly equivalent to using battery power only. This second region defined by the battery only
sector will be labeled region two. The third region is the area on the graph to the right of the
battery only sector. Within region three, power system performance is better than a battery only
power system. For a hybrid system to perform better than a battery only system, points on a
configuration line should lie within region three. Note that the boundaries of region 2 are not a
fixed ratio of flight time over weight, instead there is a range of ratios along the boundary. Along
the boundary between region 1 and 2, the flight time to weight ratios range from 46 min/kg to 67
min/kg. Along the boundary between regions 2 and 3, the flight time to weight ratio ranges from
63 min/kg to 84 min/kg.
106
Interpreting the results of the power system weight versus flight time graph of figure 57
develops several discussion points. For the battery only configuration, power system weight is the
lightest but flight time is restricted to under 2 hours when using only one battery. The flight time to
weight ratio for the three batteries (6 Ah, 11 Ah, 17 Ah) are 49.1, 71.9, and 77.9 min/kg
respectively. The trend shows that larger battery capacities yields better ratios. With the H-500
fuel cell only configuration, for the three storage tanks (Ninja 50 = 166L, Ninja 68 = 225.8 L, Ninja
90 = 298.9L), the flight time to weight ratios are 41.1, 53.5, and 67.4 min/kg respectively. Better
ratios are obtained with larger hydrogen storage tanks. As seen on the graph of figure 57, the H-
500 fuel cell only configuration line lies almost completely in region 1. In other words, using the H-
500 fuel cell only configuration with the Ninja 68 tank yields worse performance than loading the
UAV with six 6 Ah batteries. Using the H-500 fuel cell with the Ninja 90 tanks yields worse
performance than loading the UAV with three 17 Ah batteries but slightly better performance than
eight 6 Ah batteries. Next, for the three H-100 hybrid configurations, much of the configuration
lines lie within region 2. That is, using the H-100 fuel cell with Ninja 50 or 68 tanks yields flight
time to weight ratios similar to battery only configurations. Only when using the Ninja 90 tanks, do
the H-100 configuration lines extend into region 3. The H-100 hybrid configurations using the
Ninja 90 tank give flight time to weight ratios of 90.1, 92.0, and 91.8 min/kg for the 6 Ah, 11 Ah,
and 17 Ah batteries respectively. One would like to see better performance from hybrid power
configurations, but the H-100 fuel cell is a general purpose fuel cell, not weight optimized for
aviation applications. Next, for the AST01-01 hybrid configurations, all maximum flight time points
lie in region 3. Any hybrid power system configuration using the AST01-01 fuel cell performs
better than battery only systems. The flight time to weight ratios of the AST01-01 hybrid
configurations using the Ninja 90 tank are 129.9, 128.8, and 123.5 min/kg for the 6 Ah, 11 Ah,
and 17 Ah respectively. The AST01-01 fuel cell is weight optimized for aviation applications and
the weight savings significantly improve performance to be superior to battery only solutions. The
107
CHAPTER 5
Using the Simscape language to model the battery, fuel cell, and electrical loads proved
to be helpful at both the component level and the system level. The main advantage of the
Simscape physical modeling approach over the conventional Simulink approach was the ability to
partition the model into independent physical component blocks. The battery, fuel cell, motor-
propeller, and BOP models were developed separately from the top level power systems model.
Each component was tested and verified in isolation which simplified the modeling task. When
the top level hybrid power system model was assembled, there were no concerns about changing
The battery model is based on a commercially available LiPo 4S 8Ah battery and uses
the Tremblay battery model reported in literature (Tremblay & Dessaint, 2009). Implementation of
the model is realized as a Simscape text file which is listed in appendix A. To verify the battery
model and to adjust the model parameters, an adjustable constant current load is connected to
the battery. Parameters for the battery model are then adjusted to match the discharge curves of
the actual LiPo 4S battery over a range of discharge rates. The battery model verification result is
presented as a group of battery discharge curves plotted in figure 17. To check the validity of the
battery model in a second way, the battery discharge curve for a 0.5 C discharge rate is
compared against the Sim Power Systems lithium ion battery model. Although the two curves are
not a perfect match, they are in close agreement. The combined plot of the Tremblay battery
model and the Sim Power Systems lithium ion battery model 0.5 C discharge curves are shown in
figure 18.
The fuel cell model is based upon the Horizon H-100 fuel cell and uses the electrical
circuit equivalent model shown in figure 24 (b). Alternative mathematical expressions for the
Nernst potential, activation polarization, and ohmic polarization are presented by Lee in literature
and are used in this fuel cell model (Lee & Wang, 2007, Qingshan et al., 2008). The
implementation of the fuel cell model is a hierarchical Simscape model where each
108
subcomponent consists of a text file and the top level fuel cell component is another text file
which assembles the subcomponents. Listings of the subcomponent models and the fuel cell
component model are listed in appendices D, F, H, I, J, K, and L. To test the fuel cell model, an
adjustable constant current load is connected to the fuel cell. During a test simulation, the load
current is varied and the fuel cell voltage and current values are recorded (figure 29). The fuel cell
voltage is then plotted against the current to obtain the polarization curve. In addition, fuel cell
power is also plotted against current yielding the fuel cell power curve. Both plots are presented
in figure 30. The fuel cell model parameters are adjusted so that the model polarization curve
matches the polarization curve reported for the H-100 fuel cell. Two key points on the polarization
curves are the nominal operating point at (8.3 A, 12 V) and the open circuit point at (0A, 19V).
The power curve peaks at (8.3 A, 100 W) in agreement with the H-100 fuel cell power curve
(Horizon, 2013). The dynamic response of the fuel cell is also simulated for a range of double
The electrical load consists of two components, the motor-propeller model and the BOP
model. The BOP model is implemented as a Simscape graphical file and consists of a small DC-
to-DC converter and a programmable resistor load (figure 33). A signal builder block controls the
load resistance and simulates a 10 W BOP load with short intervals of 17 W load to represent the
fuel cell fan load. The motor-propeller model is also a Simscape graphical file (figure 35). The
motor-propeller electrical load is based upon data obtained with the TTB and motor manufacturer
data. A test simulation varies the throttle input and records the output electrical and mechanical
power. The end result of the test simulation are two graphs, propeller speed and power versus
The HPMC concept is developed with the CAOC described in section 4.1. Instead of
using a DC-to-DC converter in the main power path, a MOSFET is used to regulate both battery
and fuel cell current. A setpoint voltage determines a fuel cell threshold current for which any load
current demands above the threshold are supplied by the battery. The setpoint may also be set
below the cruise load current value so that the battery contributes to cruise power. The HPMC is
implemented as a Simscape graphical file shown in figure 40. Controlling the CAOC setpoint
109
voltage is a Stateflow block which simulates the function of a microprocessor. Functions of the
Stateflow block are discussed in section 4.2. Simulation tests of the HPMC show the CAOC
functioning as expected (figures 45 and 46). To verify that the fuel cell and battery are operating
within their safe zones, figure 45 shows the fuel cell current is maintained under its maximum
operating current and figure 47 shows the lowest battery SOC is kept above 35%.
The complete hybrid power system is assembled as a graphical Simscape file and is
displayed in figure 43. The flight profile is the input to the power system model which simulates a
typical six hour flight. The flight profile consists of three signals, the throttle command, the BOP
load, and a logic control signal (BAS signal). A graph of the six hour flight profile is shown in
figure 44. The simulation results presented in section 4.4 show that 366.5 L of hydrogen are
consumed without current assist and 327 L of hydrogen are used when current assist is active.
Battery SOC and fuel cell current are maintained in their safe zone of operation as shown in
The battery charger model is developed in section 3.3 and the Simscape graphical file is
displayed in figure 19. The battery charger model uses a small DC-to-DC converter to create a 24
volt supply that is combined with a programmable constant current circuit. To test the battery
charger, it is connected to the battery model and a simulation is run where the battery starts out
at 40% SOC and is charged at a 0.1 C rate up to 96% SOC. Results of this test simulation show
the battery SOC increasing linearly in figure 21, the battery voltage during recharge in figure 22,
and the constant charge current in figure 23. Inflight battery recharge simulations are then carried
out in section 4.5. Additional logic is added to the Stateflow block within the HPMC block to
control the charger (figure 52). Inflight battery recharge is simulated and the results displayed in
figure 55 show the battery SOC increasing during the flight. The Stateflow logic that controls the
battery charger sets a battery SOC upper limit at 90%. During simulation, this upper limit is
reached and figure 55 shows the charge cycle ending at the 90% upper limit.
To derive an algebraic expression to calculate flight time for different batteries and
hydrogen storage tanks, the input flight profile is divided into three sections, takeoff, cruise, and
landing. Noting that the throttle command is roughly constant during cruise, this attribute is
110
leveraged to determine flight time. The energy consumed during takeoff and landing are
subtracted out and the flight time is determined by the constant energy consumption during cruise
plus the time elapsed during takeoff and landing. Next, the usable energy from the stored
hydrogen and battery are combined and used to determine potential flight time. The resulting
flight time equation is listed as equation 52. Next, a group of commercially available batteries and
compressed air tanks are selected and their weights and capacities are listed in tables 10 and 12
respectively. The flight time equation is then used to approximate potential flight times for nine
power system configurations. The first configuration consists of a battery only configuration and
flight times are calculated for 3 battery sizes. A second configuration consists of a Horizon H-500
fuel cell only and flight times are calculated for three hydrogen storage tank sizes. Six hybrid
power configurations consist of the Horizon H-100 or AST01-01 fuel cell and a 6 Ah, 11 Ah, or 17
Ah battery. Flight times are calculated for all six hybrid configurations. The last configuration is a
conventional hybrid power configuration using the H-100 fuel cell, 6 Ah battery and a DC-to-DC
converter. This last configuration serves as a baseline for comparison of the other hybrid
configurations. The results are displayed as a graph of power system weight versus flight time
shown in figure 57. The increased ratio of flight time to weight for the AST01-01 hybrid
configurations over the H-100 hybrid configurations demonstrate the importance of minimizing
weight in UAV power systems. Results from figure 57 also demonstrate the importance of
hydrogen storage in UAV applications. The more hydrogen that can be stored onboard, the
111
5.2 Future Research
The series regulator component inside the CAOC is the MOSFET Q1. During periods of
high battery current, the power dissipated by Q1 can reach 100 W. Referring to figure 48, the
takeoff portion of the flight profile is where Q1 dissipates the most power. In order to keep Q1
internal junction temperature below its rated maximum (usually 150oC), a heat sink must be used.
A future research opportunity is to design a heat sink for Q1 and evaluate Q1 junction
temperature during flight with extra attention focused on the takeoff period. A possible location for
the heat sink is to place it on the upper surface of the UAV fuselage; the outer surface of the heat
sink being flush with the outer surface of the fuselage and the inner surface of the heat sink
mated to transistor Q1. UAV motion through the air creates an airflow over the heat sink which
enhances the cooling capacity of the heat sink. During takeoff, the UAV starts at zero velocity and
gains speed as it proceeds down the runway. This period of time is where the heat sink must
dissipate the most power but the airflow over the heat sink is minimum. A finite element analysis
(FEA) thermal analysis of the power dissipation profile (figure 48) for Q1 combined with a minimal
or increasing airflow over the heatsink can simulate Q1 junction temperature during the takeoff
period. Another scenario where heat sink analysis may be necessary is during cruise when
current assist is active. Although the power dissipated by Q1 is much lower, around 1.3 W, it is
To improve hydrogen storage capacity onboard the UAV, liquid hydrogen storage can be
used in place of compressed hydrogen storage tanks. Liquid hydrogen (LH2) is an attractive
alternative to compressed hydrogen storage due to its high volumetric density of 70.8 g/L
(Khandelwal et al., 2013). When compared to the volumetric energy density of hydrogen gas at
STP (9.829 MJ/L), LH2 has 864.3 times the volumetric energy density (8495 MJ/L). The clear
advantage of liquid hydrogen is balanced by the technical difficulties of storage and the highly
energy intensive process of liquefying hydrogen. The source of complication to store liquid
hydrogen is the inherent temperature and pressure range at which liquid hydrogen exists. Typical
LH2 storage tanks are kept at a temperature of 20.5oK and at a pressure of 1 to 3 bars
112
(Khandelwal et al., 2013). To maintain its liquid state, the LH2 must be kept below the critical point
of 33oK and 12.9 bar and above its triple point of 13.8oK and 0.0719 bar (“Properties of
Hydrogen”, 2016).
A cryogenic pressure vessel named a cryostat is a temperature controlled storage tank used to
contain LH2. Designed as a dual shell system, the inner and outer metallic shells are separated
by a sandwiched layer of insulation. Careful thermal design must account for thermal conduction,
convection, and radiation in order to maintain the extremely low storage temperature. The inner
shell contains the cryogenic hydrogen and uses a multi-layered material as an insulator wrap
(Tzimas, 2003). The insulation wrap consists of a reflective foil on the inner surface which reflects
radiative thermal transfer, and then an alternating stack of thin insulating material such as a glass
fiber and thin layers of aluminum (Khandelwal et al., 2013). Mechanical spacers are placed
between the inner and outer shells to maintain a controlled gap between the two vessels. The
gap is then evacuated to a near perfect vacuum to minimize thermal convection, thus maximizing
the thermal insulation. Despite the extreme measures to create a perfectly insulating pressure
vessel, there is still heat leakage. Additionally, the orthohydrogen content is never zero in
cryogenic hydrogen, and when it changes to parahydrogen, heat is generated internally. Due to
heat effects, there is always a certain amount of LH2 that “boils off”. The boil off rate will
determine the effective storage time of a cryostat. An acceptable boil off rate for aviation
113
applications is 0.1% of LH2 by weight per hour (Khandelwal et al., 2013). An example LH2 storage
system for an automotive application has a gravimetric density of 14.2% (Larminie & Dicks,
2003). Despite the complicated implementation of a liquid hydrogen storage system, the NRL
team demonstrated its use in a UAV application with the 2013 flight of the Ion Tiger (“Ion Tiger
Fuel Cell Powered UAV”, 2016). A second notable example of a UAV powered by LH2 fuel is the
Global Observer built by AeroVironment. Designed as an ISR platform, the Global Observer flew
5.3 Conclusion
Interest in sUAS has been on the rise as the costs decrease and civilian applications and
opportunities expand. Within the aerospace community, the sUAS sector has been recognized as
the most dynamic new market with sales expected to exceed 8 billion US dollars by 2018
(Manufacturing Close-Up, 2015). Continued research in the area of electric powered UAVs has
made steady progress, increasing UAV capabilities such as flight endurance. To bring UAV flight
endurance to the next level, a key and disruptive technology is the use of fuel cells as an electric
power source. Integrating fuel cell technology into sUAS continues to challenge engineering
practice as fuel cell weight and limited power delivery are barriers to overcome. Combining
conventional lithium ion batteries with PEMFCs into a hybrid power system has the advantage of
minimizing fuel cell weight while still providing high power when needed. Many APM schemes
reported in literature optimize hybrid power system performance through the use of DC-to-DC
converters between power sources. The issue with this approach is that the DC-to-DC converters
are in the main power path and therefore require high power ratings. High power DC-to-DC
converters add significant weight to the power system. This study adds to APM literature with a
new design approach that uses a transistor in place of the DC-to-DC converter. The transistor
weight is much lower than the weight of a high power DC-to-DC converter thus reducing weight of
the power system. In addition, significant power is dissipated by the transistor only during high
power delivery from the battery. During the majority of the flight, power dissipated by the
transistor can drop to zero. Over the duration of the flight, power system efficiency of the
114
transistor APM approach exceeds 98% which is greater than typical efficiencies achieved with
DC-to-DC converters.
The Simscape model of the UAV power system developed in this study demonstrates
several points. First, it verifies the functionality of the transistor-based APM approach. Second, it
implements the Tremblay battery model as a Simscape model and verifies model behavior in a
larger system. Similarly, the Lee fuel cell model is implemented as a Simscape model and verified
in system level simulations. Third, the flight simulations determine the amount of hydrogen
required for the ASU Composite UAV to achieve a six hour flight time. Fourth, the flight
simulations with current assist activated demonstrate that battery power can be used during
cruise flight to enhance flight endurance or reduce the amount of hydrogen necessary to achieve
a flight time goal. Fifth, the UAV power system model demonstrates inflight battery recharge
which can be useful for certain scenarios. Sixth, implementing the CAOC in the HPMC
demonstrates that monitoring fuel cell and battery voltage and current allows a microprocessor to
implement an observer algorithm that can closely track battery SOC, fuel cell efficiency, and
hydrogen consumption. The final point demonstrated by the UAV power system simulations are
the potential flight times that can be expected for various hybrid power system configurations.
Overall, the UAV hybrid power system model proved useful in many areas.
115
REFERENCES
AXI Motor Applications 4120/14 (2015), Retrieved November 24, 2015, from
http://www.modelmotors.cz:
http://www.modelmotors.cz/index.php?page=61&product=4120&serie=14&line=GOLD.
Bard, A., Faulkner, L. (2001). Electrochemical methods fundamentals and applications. Wiley.
Bradley, T. (2008). Modeling, design and energy management of fuel cell systems for aircraft.
Georgia Institute of Technology Dissertation.
Buasri, P., & Salameh, Z. (June, 2006). An electrical circuit model for a proton exchange
membrane fuel cell (PEMFC). Power Engineering Society General Meeting, IEEE 2006.
Current Sharing in Power Arrays (n.d.). Maxi, Mini, Micro Design Guide, rev. 4.9, pp. 20 – 24,
Retrieved September 17, 2016, from www.vicorpower.com
http://www.vicorpower.com/documents/applications_manual/05current_sharing.pdf.
DOD dictionary (2007). Dictionary of Military and Associated Terms. JP 1-02, 569.
DOE FCTO (2016, May). Hydrogen fuel cells for small unmanned air vehicles. Retrieved
September 11, 2016 from http://energy.gov:
http://energy.gov/sites/prod/files/2016/05/f32/fcto_webinarslides_h2_fc_small_unmanned_air_veh
icles_052616.pdf.
EnergyOr fuel cell powered UAV reaches 10 h flight endurance (2011, September). Fuel Cell
Bulletin, 2011, pp. 4-5.
FAUCON H2 UAV image (2016). Retrieved September 16, 2016, from http://energyor.com:
http://energyor.com/static/template/eo/images/products/h2-endurance/MB-99-l.jpg.
Fuel cell/battery hybrid UAV takes off in Taiwan (2010, June). Fuel Cell Bulletin, 2010, pp. 4-5.
Fuentes, G. (2005, April 4) DARPA Test-Flies 7-Ounce UAV. Defense News. p. 26.
Furrutter, M., Meyer, J. (2009, September). Small fuel cell powering an unmanned aerial vehicle.
AFRICON 2009, pp. 1-6.
General Atomics MQ-1 Predator (June 9, 2016). Retrieved June 18, 2016 from
https://en.wikipedia.org:
https://en.wikipedia.org/wiki/General_Atomics_MQ-1_Predator.
116
Gibbs, Y., (July 30, 2015). NASA Armstrong fact sheet: Altair. Retrieved from
http://www.nasa.gov:
http://www.nasa.gov/centers/armstrong/news/FactSheets/FS-073-DFRC.html.
Grey-faced Buzzard UAV image (2016), Retrieved June 16, 2016, from http://cdn.ubergizmo.com:
http://cdn.ubergizmo.com/photos/2010/5/asia-hybrid-uav.jpg.
Gundlach, J. (2014). Designing unmanned aircraft systems, a comprehensive approach, (pp. 14-
22). AIAA.
Horizon (August, 2013). H-100 fuel cell stack user manual. Retrieved February 10, 2016, from
http://media.wix.com:
http://media.wix.com/ugd/047f54_bc938ab51966e498e9919dfebb88e09d.pdf
Hydrogen phase diagram image (2016). Retrieved September 23, 2016, from
https://hydropole.ch:
https://hydropole.ch/wp-content/uploads/2014/12/1372142909_phasediag.gif.
Interstate TDR (May 18, 2016). Retrieved June 17, 2016, from https://en.wikipedia.org:
https://en.wikipedia.org/wiki/Interstate_TDR.
Ion Tiger Fuel Cell Powered UAV (2016). Retrieved September 10, 2016, from
http://www.nrl.navy.mil:
http://www.nrl.navy.mil/lasr/content/ion-tiger-fuel-cell-powered-uav.
Ion Tiger UAV image. Retrieved September 11, 2016, from http://www.nrl.navy.mil:
http://www.nrl.navy.mil/PressReleases/2010/image3_3-10r.jpg.
Jefferson, M. (2002, December 17) DARPA to seek long endurance with ‘Hornet’ micro air
vehicle. Aerospace Daily.
Jiang, Z., Gao, L., Blackwelder, M., & Dougal, R. (December, 2003). Design and experimental
tests of control strategies for active hybrid fuel cell/battery power sources. Journal of Power
Sources, 2003 (130), pp. 173 – 171.
Jiang, Z. Gao, L., & Dougal, R. (June, 2007). Adaptive control strategy for active power sharing in
hybrid fuel cell/battery power sources. IEEE Transactions on Energy Conversion, 2007 (22), pp.
507 – 515.
Jones, P., Lakeman, J., Mepsted, G., & Moore, J. (December, 1998). A hybrid power source for
pulse power applications. Journal of Power Sources, 1998 (80), pp. 242 – 247.
Khandelwal, B., Karakurt, A., Sekaran, P., Sethi, V., & Singh, R. (January, 2013). Hydrogen
powered aircraft : The future of air transport. Progress in Aerospace Sciences, (60) 2013.
Keane, J., & Carr, S. (2013). A brief history of early unmanned aircraft. Johns Hopkins APL
Technical Digest, 32(1), 558-571. Retrieved June 17, 2016, from http://www.jhuapl.edu/:
http://www.jhuapl.edu/techdigest/TD/td3203/32_03-Keane.pdf.
Krauch, T., Berg, S., Beckstead, A., Wilkinson, D., Coronel, M., Kastelic, G., Padron, N., & Vo, D.
(December, 2012). Design, construction, and testing of a composite multi-role UAV load bearing
aircraft. Arizona State University Project Report.
117
Langley Aerodrome Number 6 (n.d.). Retrieved June 17, 2016, from http://airandspace.si.edu:
http://airandspace.si.edu/collections/artifact.cfm?object=nasm_A19050002000.
Larminie, J., & Dicks, A. (2003). Fuel cell systems explained, second edition. West Sussex: J.
Wiley & Sons.
Lee, B. (2013). TDR-1 First Operational US Navy Drone Successful in Combat in 1944. Retrieved
June 18, 2016, from http://www.nnapprentice.com/
http://www.nnapprentice.com/alumni/letter/TDR_1.pdf.
Lee, B., & Park, P. (October, 2014). Active power management system for an unmanned aerial
vehicle powered by solar cells, fuel cell, and batteries. IEEE Transactions on Aerospace and
Electronic Systems, 2014 (50), pp. 3167 – 3177.
Lee, D., & Wang, L. (June, 2007). Dynamic and steady-state performance of PEM fuel cells under
various loading conditions. IEEE Power Engineering Society General Meeting.
Lockheed Martin ruggedized UAS uses AMI fuel cell power (September, 2011). Fuel Cell Bulletin,
2011 (9), p. 4.
Linden, D., & Reddy, T. (2002). Handbook of batteries, third cdition. McGraw-Hill.
Lycoming O-435 (September 25, 2015). Retrieved June 17, 2016, from https://en.wikipedia.org:
https://en.wikipedia.org/wiki/Lycoming_O-435.
Manufacturing Close-Up (Jan. 8, 2015). ABI research: Commercial sector is sweet spot for sUAS
market. Close-Up Media, Inc.
MaxAmps LiPo 4S 8Ah (September, 2016). Retrieved September 10, 2016, from
www.maxamps.com:
https://www.maxamps.com/lipo-8000-4s-14-8v-dual-core-battery-pack.
MaxAmps LiPo 4S Family (October, 2016). Retrieved October 9, 2016, from www.maxamps.com:
https://www.maxamps.com/products/4s-14-8v-lipos.
McConnell, V. (2007, December) Military UAVs claiming the skies with fuel cell power. Fuel Cell
Bulletin. pp. 12-15.
Moffit (2010). A methodology for the validated design space exploration of fuel cell powered
unmanned aerial vehicles. Georgia Institute of Technology Dissertation.
Monaco, E. (2013). Optimization of flight time using fuel cell on unmanned aerial vehicles.
Arizona State University Project Report.
MQ-1B Predator (September 23, 2015). Retrieved June 18, 2016, from http://www.af.mil/:
118
http://www.af.mil/AboutUs/FactSheets/Display/tabid/224/Article/104469/mq-1b-predator.aspx.
MQ-1 Predator/MQ-9 Reaper (2016). Retrieved June 17, 2016 from http://www.bga-
aeroweb.com:
http://www.bga-aeroweb.com/Defense/MQ-1-Predator-MQ-9-Reaper.html.
NASA (August, 2006). Earth observations and the role of UAVs: A capabilities assessment,
version 1.1, (p. 12). Retrieved June 16, 2016, from http://www.nasa.gov:
http://www.nasa.gov/centers/dryden/research/civuav/index.html.
Ninja Pro v2 90/4500 tank image (2016). Retrieved September 22, 2016, from
www.ninjapaintball.com:
http://www.ninjapaintball.com/tanks?lightbox=image_1dk3.
Ninja Pro v2 90/4500 specs (2016). Retrieved September 22, 2015, from www.evike.com:
http://www.evike.com/products/50046/.
Northrop Grumman Facts (May, 2008). Retrieved June 17, 2016 from
http://www.northropgrumman.com:
http://www.northropgrumman.com/capabilities/rq4block20globalhawk/documents/hale_factsheet.
pdf.
Northrop Grumman RQ-4 Global Hawk (June 11, 2016). Retrieved June 17, 2016, from
https://en.wikipedia.org:
https://en.wikipedia.org/wiki/Northrop_Grumman_RQ-4_Global_Hawk.
NRL Fact Book (2014). Retrieved September 11, 2016, from http://www.nrl.navy.mil:
http://www.nrl.navy.mil/content_images/2014_FactBook_cleansed.pdf.
Piper Cub UAV image (2016). Retrieved September 16, 2016, from www.google.com:
https://www.google.com/search?q=University+of+Johannesburg+Fuel+Cell+UAV+image&biw=11
73&bih=629&tbm=isch&imgil=CrGj8aTkfFlw_M%253A%253BjpZaRc4C3jXF5M%253Bhttp%252
53A%25252F%25252Fwww.unmannedsystemstechnology.com%25252F2015%25252F08%2525
2Famerican-university-of-sharjah-conducts-first-hydrogen-fuel-cell-powered-uav-flight-in-
uae%25252F&source=iu&pf=m&fir=CrGj8aTkfFlw_M%253A%252CjpZaRc4C3jXF5M%252C_&u
sg=__JWg-
_PRowNXbAPMgIbBGdR5D59g%3D&ved=0ahUKEwiojv2a5pTPAhVD7GMKHYMCBSQQyjcINQ
&ei=tF_cV6jkO8PYjwODhZSgAg#imgrc=CrGj8aTkfFlw_M%3A.
Qingshan, X., Nianchun, W., Ichiyanagi, K. & Yukita, K. (April, 2008). PEM fuel cell modeling and
parameter influences of performance evaluation. Electric Utility Deregulation and Restructuring
and Power Technologies, Third International Conference.
Revankar, S., & Majumdar, P. (2014). Fuel cells principles, design, and analysis. CRC Press.
119
Sadraey, M. (2013). Aircraft design, a systems engineering approach. Wiley, p.127.
Shepherd, C. (1965, July). Design of primary and secondary cells II. An equation describing
battery discharge. Journal of Electrochemical Society, 1965 (112), pp.657-664.
Spider Lion UAV image (2016). Retrieved September 10, 2016, from https://www.google.com:
https://www.google.com/search?q=Spider+Lion+UAV+image&biw=1261&bih=686&tbm=isch&img
il=Zmf3bIJ5rti8qM%253A%253B45l3yXl_LjbCGM%253Bhttp%25253A%25252F%25252Fwww.d
esignation-systems.net%25252Fdusrm%25252Fapp4%25252Fspider-
lion.html&source=iu&pf=m&fir=Zmf3bIJ5rti8qM%253A%252C45l3yXl_LjbCGM%252C_&usg=__2
F0scuC-t9v5fQsNBKu2edawbgQ%3D&dpr=1.2&ved=0ahUKEwiwqe2w1ofPAhVN-
2MKHeUpANEQyjcIOg&ei=R37VV_DpEc32jwPl04CIDQ#imgrc=Zmf3bIJ5rti8qM%3A.
Stalker UAV image (2016). Retrieved September 16, 2016, from http://www.lockheedmartin.com:
http://www.lockheedmartin.com/us/products/stalker-uas.html.
Suh, K. (2006). Modeling, analysis, and control of fuel cell hybrid power systems. University of
Michigan Dissertation.
Swider-Lyons, K., MacKrell, J., Rodgers, J., Page, G., Schuette, M., & Stroman, R. (September,
2011). Hydrogen fuel cell propulsion for long endurance small UAVs. AIAA centennial of Naval
Aviation Forum “100 Years of Achievement and Progress”, AIAA 2011-6975.
Swider-Lyons, K. (May, 2016). Hydrogen fuel cells for small unmanned air vehicles. DOE
presentation May 26, 2016, Retrieved September 23, 2016, from http://energy.gov:
http://energy.gov/sites/prod/files/2016/05/f32/fcto_webinarslides_h2_fc_small_unmanned_air_veh
icles_052616.pdf.
Targets for Onboard Hydrogen Storage Systems for Light-Duty Vehicles (September, 2009).
Freedom Car Fuel Partnership, DOE.
Tawfik, H., Hung, Y., & Mahajan, D. (September, 2006). Metal bipolar plates for PEM fuel cell – A
review. Journal of Power Sources, 2006 (doi:10.1016/j.jpowsour.2006.09.088).
TPE331-10 Turboprop Engine (April, 2006). Retrieved June 17, 2016, from
http://www51.honeywell.com:
http://www51.honeywell.com/aero/common/documents/myaerospacecatalog-
documents/BA_brochures-documents/TPE331.10.pdf.
Tremblay, O., & Dessaint, L. (May, 2009). Experimental validation of a battery dynamic model for
EV applications. World Electric Vehicle Journal, 2009 (3), pp. 289-299.
Tsang, K., Sun, L., & Chan, W. (June, 2010). Identification and modelling of lithium ion battery.
Energy Conversion and Management, 2010 (51), pp. 2857-2862.
120
Tzimas, E., Filiou, C., Peteves, S., & Veyret, J. (2003). Hydrogen storage: State-of-the-art and
future perspective. Institute for Energy, Petten, the Netherlands. Retrieved September 23, 2016,
from www.Linde.com:
http://publications.jrc.ec.europa.eu/repository/bitstream/111111111/6013/1/EUR%2020995%20E
N.pdf.
USAF contract for Protonex UAV power (March, 2006). Fuel Cell Bulletin, p. 3.
Zhang, J., Li J., Li, Y., Zhao, Y. (2014). Hydrogen generation, storage, and utilization. Wiley, pp.
75-80.
121
APPENDIX A
122
component LiPo_4S
% Battery model of LiPo 4S, 8 Ah
nodes
p = foundation.electrical.electrical; % +:right
n = foundation.electrical.electrical; % -:right
end
outputs
SOC = { 0, '1'}; % SOC:left
BatteryVoltage = { 0, 'V' }; %V:left
BatteryCurrent = { 0, 'A' }; %I:left
end
variables(Access = protected)
v = {0, 'V'}; % Across variable, voltage
i = {0, 'A'}; % Through variable, current
CT = {0, 'c'}; % Charge Transfer, coulombs
end
parameters
OCV = { 15.6, 'V' }; % Battery Open Circuit Voltage (volts)
Q = { 8, 'A*hr' }; % Battery Capacity (Amp-hours)
IntRes = { 0.005, 'Ohm' }; % Internal Series Resistance (Ohms)
InitialSOC = { 100.0, '1' }; % Initial State of Charge (percentage)
K = { 0.0075, 'V/(A*hr)'}; % Polarization Constant (V/(Amp-hours))
Kr = {0.0064, 'Ohm'}; % Polarization Resistance (Ohm)
A = { 1.2, 'V' }; % Exponential Zone amplitude (V)
B = { 2.0, '1/(A*hr)' }; % Exponential Zone time constant inverse (1/(Amp-hours))
end
function setup
if OCV <= 0
error('OCV must be greater than zero');
end
if Q <= 0
error('Battery capacity must be greater than zero');
end
% From the Initial SOC and the Capacity (Q), determine
% the starting value for CT (charge transferred)
CT = Q*{3600,'c/(A*hr)'}*(1 - (InitialSOC/100));
end
branches
i: n.i -> p.i; % define current flow (source: neg to pos)
end
equations
v == p.v - n.v; % Voltage across battery (pos to neg)
let
ohmic = i * IntRes;
ExpZone = A*exp(-B*(CT));
PolVolt = K*(Q/(Q - CT)) * CT;
PolResDischarge = Kr*(Q/(Q - CT)) * i;
PolResCharge = Kr*(Q/(CT - Q)) * i;
123
in % Battery equation
if i >= 0
v == OCV + ExpZone - PolVolt - ohmic - PolResDischarge;
else
v == OCV + ExpZone - PolVolt - ohmic - PolResCharge;
end
end
i == CT.der; % Current is derivative of Charge Transferred
SOC == 100.0*(1 - (CT/Q)); % State of Charge
BatteryVoltage == v; % export the battery voltage
BatteryCurrent == i; % export the battery current
end
end
124
APPENDIX B
125
% Discharge_4S_curves.m
% LiPo_4S battery discharge curve graph
%
% Plot the discharge curves for the LiPo_4S, 8 Ah battery for
% 0.5C, 1.0C, 2.0C, 4.0C, 8.0C 16.0 C
DR = [0.5 1.0 2.0 4.0 8.0 16.0]; % battery discharge rate vector
tmax = zeros(1,length(DR)); % array of maximum simulation times
dlc = [0.0 0.1 0.9; 0.7 0.4 0.2; 0.1 0.9 0.1; 0.4 0.1 0.1; 0.9 0.0 0.0; ...
0.0 0.0 0.0];
BatCap = 8.0; % battery capacity in Amp-hours
DC = BatCap * DR; % discharge currents
for j = 1:length(DR)
csp = 1.7 + 2*DC(j); % current setpoint
tmax(j) = (BatCap / DR(j))*3600; % stop time for simulation
sim('LiPo_Discharge_01', 'StopTime', num2str(tmax(j))); % run the model
xvals = simlog.C2Ah.O.series.values;
yvals = simlog.Battery_model_of_LiPo_4S_8_Ah.BatteryVoltage.series.values;
plot(xvals, yvals, 'LineWidth', 2, 'Color', [dlc(j,1) dlc(j,2) dlc(j,3)]);
if j == 1
hold on;
end
end
hold off;
title('LiPo 4S, 8 Ah, Battery Discharge Curves');
ylabel('Battery Voltage (V)');
xlabel('Charge (Ah)');
legend('0.5 C', '1.0 C', '2.0 C', '4.0 C', '8.0 C', '16.0 C', ...
'Location', 'southwest');
text(0.1, 6.3, 'Legend: Discharge Rate');
set(gca, 'FontSize', 14);
grid;
set(gcf, 'color', 'w');
set(findall(gcf, 'type', 'text'), 'FontSize', 18);
126
APPENDIX C
127
Figure 59. Battery Discharge Comparison Model
128
APPENDIX D
129
component Enernst
% Nernst Voltage
inputs
H2_pressure = { 1, 'atm' }; % H2p:bottom
Air_pressure = { 1, 'atm' }; % Airp:bottom
end
outputs
Vrev = { 0, 'V' }; % Vr:top
end
parameters
T = { 65 + 273.15, 'K' }; % cell temperature (K)
noc = { 12, '1' }; % number of cells
end
parameters(Hidden = true)
Tref = { 25 + 273.15, 'K' }; % reference temperature
end
variables(Access = protected)
v = {0, 'V'}; % Across variable, voltage
i = {0, 'A'}; % Through variable, current
end
function setup
if T < {(60 + 273.15), 'K'} || T > {(100 + 273.15), 'K'}
error('Temperature must be between 60 and 100 Celsius');
end
if noc < 1
error('number of cells must be greater than one');
end
Tref = { 25 + 273.15, 'K' };
end
branches
i: p.i -> n.i; % define current flow (source: neg to pos)
end
equations
let
Eo = {1.229,'V'} - {0.85e-3,'V/K'}*(T - Tref) + ...
{4.31e-5,'V/K'}*T*(log(H2_pressure * {1,'1/atm'}) + ...
0.5*log(Air_pressure * {0.2095,'1/atm'}));
in
v == 0.8 * noc * Eo;
end
v == p.v - n.v; % Voltage across battery (pos to neg)
130
Vrev == v;
end
end
131
APPENDIX E
132
% pe3D_Lee_Nernst.m
% Percentage error of Lee approximation of Nernst potential
% x-axis: temperature (degree Celsius), range: 25 - 200
% y-axis: hydrogen partial pressure (atm.), range: 0.5 - 3
% z-axis: ((Estd - Elee)/Estd)*100%
% constants
F = 98485; % Faraday's constant (Coulombs / mole)
R = 8.314; % Universal gas constant (J/(mole*K))
% color matrix
cm = [5 65 0; 255 69 0; 255 0 0; 218 165 32; 189 183 140;
154 205 50; 127 255 0; 0 100 0; 102 205 160; 32 178 170;
95 158 160; 0 0 205; 0 255 250; 128 0 0; 128 0 128;
0 128 128; 180 180 192; 178 34 34; 255 215 3; 250 140 5];
% normalize color matrix
cm = (1/255)*cm;
% define a function handle for interpolated data from Gibb's table
fh_gfe = @(g)interp1([25 80 100 200],[-237.2 -226.1 -225.2 -220.4], g);
% pre-allocate memory
Elee = zeros(20,20); % Lee approximation of Nernst potential
Estd = zeros(20,20); % Standard calculation of Nernst potential
term_1 = zeros(20,20); % first term in standard calculation
term_2 = zeros(20,20); % second term in standard calculation
zero_plane = zeros(7,20); % zero reference plane
z = zeros(20,20); % pre-allocate percent error matrix
for i = 1:20
for j = 1:20
Elee(i,j) = 1.229 - 0.85*10^-3 * (xlin(i) - 25) + ...
4.3085*10^-5 * (xlin(i) + 273.15)*(log(ylin(j)));
term_1(i,j) = ( (-1*fh_gfe(xlin(i)))/(2*F) )*1000;
term_2(i,j) = ((R*(xlin(i)+273.15))/(2*F))*(log(ylin(j)));
Estd(i,j) = term_1(i,j) + term_2(i,j);
z(i,j) = ((Estd(i,j) - Elee(i,j))/Estd(i,j))*100;
end
end
134
APPENDIX F
135
component Ractivation
% R activation
inputs
H2_pressure = { 1, 'atm' }; % H2p:left
Air_pressure = { 1, 'atm' }; % Airp:left
end
outputs
effective_Res = {0, 'Ohm'}; % eRa:right
end
variables(Access = private)
i = { 0, 'A' }; % Current
v = { 0, 'V' }; % Voltage
end
parameters
Area = { 22.25, 'cm^2' } % cell active area
T = { 65 + 273.15, 'K' }; % cell temperature (K)
noc = { 20, '1' }; % number of cells
end
parameters(Hidden = true)
z1 = { -1.81, 'V' } % zeta_1 coefficient
z2 = { 0.00286, 'V/K' } % zeta_2 coefficient
z3 = { 7.6e-5, 'V/K' } % zeta_3 coefficient
z4 = { -0.78e-4, 'V/K' } % zeta_4 coefficient
end
function setup
if Area < {0.1, 'cm^2'}
error('cell active area must be greater than 0.1 cm^2');
end
if T < {(60 + 273.15), 'K'} || T > {(100 + 273.15), 'K'}
error('Temperature must be between 60 and 100 Celsius');
end
% caculate z2
z2 = {0.00286, 'V/K'} + {0.0002,'V/K'}*log(Area*{1,'1/cm^2'});
end
branches
i : p.i -> n.i;
end
equations
let
% concentration of H2 at the catalytic interface (mole/cm^3)
% note: make CH2 unitless ('1')
136
CH2 = (H2_pressure * {1, '1/atm'}) / ...
(5.08e-6*exp(-498/(T*{1,'1/K'})));
in
let
% concentration of O2 at the catalytic interface (mole/cm^3)
% note: make CO2 unitless ('1')
CO2 = (0.2095 * Air_pressure * {1, '1/atm'}) / ...
(5.08e-6*exp(-498/(T*{1,'1/atm'})));
z2 = ({0.00286, 'V/K'} + ...
({0.0002,'V/K'}*log(Area*{1,'1/cm^2'})) + ...
({5.3e-5,'V/K'}*log(CH2)));
in
if i < {1e-29, 'A'}
v == -1*noc*((z1 + z2*T + z3*T*log(CO2) + ...
z4*T*log(1e-29)));
else
v == -1*noc*(z1 + z2*T + z3*T*log(CO2) + ...
z4*T*log(i));
end
end
end
% protect against division by zero
if i < {1e-12,'A'}
effective_Res == v / {1e-12, 'A'};
else
effective_Res == v/i;
end
v == p.v - n.v;
end
end
137
APPENDIX G
138
% cell_resistance_example.m
%
% Calculate the cell resistance for a PEMFC MEA over a temperature
% range of 40C to 90C and a current density range of 0.1 A/cm^2 to
% 1 A/cm^2. Plot results as a surface plot for each value of psi.
% Electronic resistance of cell, Rc = 0.02 ohms
% Active area, A = 20 cm^2
% Membrane thickness, t = 0.0051 cm. = 51um
Rc = 0.02; A = 20; t = 0.0051;
% create x-y grid (50 x 50)
xlin = linspace(40, 90, 50); % temperatures
ylin = linspace(0.1, 1.0, 50); % current densities
ulin = linspace(90, 40, 50);
vlin = linspace(1.0, 0.1, 50);
[x,y] = meshgrid(ulin, vlin);
% pre-allocate memory
z = zeros(50,50); % cell resistance
rho = zeros(50,50); % specific resistivity of Nafion membrane
num = zeros(50,50); % numerator
den = zeros(50,50); % denominator
% parameter psi (range 14 - 23)
psi_base = 13;
for w = 1:10
psi = psi_base + w;
for i = 1:50
for j = 1:50
Tk = xlin(i) + 273.15; % change temperature to Kelvin
cd = ylin(j); % current density
end
end
139
APPENDIX H
140
component Rohmic
% R ohmic
outputs
effective_Res = { 0, 'Ohm' }; % R:right
end
variables(Access = private)
i = { 0, 'A' }; % Current
v = { 0, 'V' }; % Voltage
end
parameters
Area = { 22.25, 'cm^2' }; % cell active area (cm^2)
T = { 65 + 273.15, 'K' }; % cell temperature (K)
l = { 0.051, 'cm' } % thickness of Nafion 112:2 (cm)
noc = { 20, '1' }; % number of cells
end
parameters(Hidden = true)
symbol_Psi = {23, '1'};
Rc = {0.04, 'Ohm'} % elecctronic cell resistance
end
function setup
if Area < {0.1, 'cm^2'}
error('cell active area must be greater than 0.1 cm^2');
end
if T < {(60 + 273.15), 'K'} || T > {(100 + 273.15), 'K'}
error('Temperature must be between 60 and 100 Celsius');
end
Rc = {0.04, 'Ohm*cm^2'} / Area;
end
branches
i : p.i -> n.i;
end
equations
let
Alpha = (((T*{1,'1/K'})/303)^2) * ...
(((i*{1,'1/A'})/(Area*{1,'1/cm^2'}))^2.5);
Beta = symbol_Psi - 0.634 - ...
3*((i*{1,'1/A'})/(Area*{1,'1/cm^2'}));
in
let
num = 36*(1 + ...
0.03*((i*{1,'1/A'})/(Area*{1,'1/cm^2'})) + ...
0.062*Alpha) * {1, 'Ohm'};
141
den = Beta*exp(4.18 * ...
(((T*{1,'1/K'}) - 303)/(T*{1,'1/K'}))) * {1, '1/cm'};
in
let
Rho = num / den; % specific resistivity (ohm*cm)
in
let
% ionic resistance of cell membrane
Rm = (Rho * l)/Area;
in
v == i*noc*(Rm + Rc);
effective_Res == noc*(Rm + Rc);
end
end
end
end
v == p.v - n.v;
end
end
142
APPENDIX I
143
component Rconcentration
% R concentration
outputs
effective_Res = { 0, 'Ohm' }; % R:right
end
variables(Access = private)
i = { 0, 'A' }; % Current
v = { 0, 'V' }; % Voltage
end
parameters
Area = { 22.25, 'cm^2' }; % cell active area (cm^2)
Jmax = { 0.43, 'A/cm^2' }; % max current density (A/cm^2)
noc = { 20, '1' }; % number of cells
B = { 0.076, 'V' }; % concentration loss coefficient
end
function setup
if Area < {0.1, 'cm^2'}
error('cell active area must be greater than 0.1 cm^2');
end
if Jmax < {0.01, 'A/cm^2'} || Jmax > {10, 'A/cm^2'}
error('Jmax must be between 0.01 A/cm^2 and 10 A/cm^2');
end
end
branches
i : p.i -> n.i;
end
equations
v == p.v - n.v;
v == (-noc*B)*log( 1 - ((i/Area)/Jmax) );
144
APPENDIX J
145
component dlCap
% double-layer Capacitor
nodes
p = foundation.electrical.electrical; % p:top
n = foundation.electrical.electrical; % n:bottom
end
outputs
Vdl = {0, 'V'}; % Vdl:bottom
Idl = {0, 'A'}; % Idl:bottom
end
variables
i = { 0, 'A' }; % Current
v = { 0, 'V' }; % Voltage
end
parameters
C = { 1.0, 'uF' }; % Capacitance value
CIV = { 0.0, 'V' }; % initial voltage
end
function setup
if C <= 0
error('Capacitance Value must be positive');
end
branches
i : p.i -> n.i; % Through variable i from node p to node n
end
equations
v == p.v - n.v; % Across variable v from p to n
i == C*v.der; % Capacitor equation
Vdl == v;
Idl == i;
end
end
146
APPENDIX K
147
component Flow
% calculate flow rate (SLPM)
inputs
I = {0, 'A'}; % I:left
end
outputs
H2_flow = {0, 'lpm'}; % H2f:right
Air_flow = {0, 'lpm'}; % Airf:right
end
parameters(Hidden = true);
F = { 96485, 'c/mol' }; % Faraday's constant
noc = { 12, '1' }; % number of cells
end
equations
% note: Simscape has an implicit conversion from liters per second
% to liters per minute. If you manually do this conversion
% as I tried, the SLPM value will be 60 times too large
% since Simscape does not recognize it.
% So, I had to comment out the Liter per second to
% Liters per minute conversion.
H2_flow == ((noc*I)/(2*F))*{24.46, 'l/mol'};%*{60,'s/min'};
Air_flow == ((noc*I)/(4*(0.2095)*F))*{24.46, 'l/mol'};%*{60,'s/min'};
end
end
148
APPENDIX L
149
component MyPEMFC_2
% PEMFC model 2
inputs
H2_pressure = {1, 'atm'}; % H2p:right
Air_pressure = {1, 'atm'}; % Airp:right
end
outputs
H2_flow = {0, 'lpm'}; % H2c:left
Air_flow = {0, 'lpm'}; % Airc:left
sVoltage = {12, 'V'}; % sV:left
sCurrent = {0, 'A'}; % sC:left
eVr = {0, 'V'}; % Vrev:left
eRa = {0, 'Ohm'}; % eRact:left
eRo = {0, 'Ohm'}; % eRohm:left
eRc = {0, 'Ohm'}; % eRcon:left
Vdl = {0, 'V'}; % Vdl:left
Idl = {0, 'A'}; % Idl:left
end
parameters
c_noc = {20, '1'}; % number of cells
c_T = { (65 + 273.15), 'K' }; % temperature (K)
c_Area = { 22.25, 'cm^2' }; % cell active area (cm^2)
c_l = { 0.051, 'cm' }; % thickness of Nafion membrane (cm)
c_B = { 0.076, 'V' }; % concentration loss coefficient (V)
c_Jmax = { 0.43, 'A/cm^2' }; % max current density (A/cm^2)
c_CapVal = { 1.0, 'uF' }; % double-layer capacitor value
c_CIV = { 0.0, 'V' }; % capacitor initial voltage
end
components(Hidden = true)
Erev = Ch3Lib.Enernst(T = c_T, noc = c_noc);
Ract = Ch3Lib.Ractivation(noc=c_noc);
RLUT = Ch3Lib.RactLUT;
Rcon = Ch3Lib.Rconcentration(Area = c_Area,Jmax = c_Jmax,...
noc = c_noc, B = c_B);
Rohm = Ch3Lib.Rohmic(Area=c_Area, T=c_T, l=c_l, noc=c_noc);
Cap = Ch3Lib.dlCap(C = c_CapVal, CIV = c_CIV);
% sensors
CSense = foundation.electrical.sensors.current;
VSense = foundation.electrical.sensors.voltage;
% calculation block (calculate flow rates)
gFlow = Ch3Lib.Flow(noc = c_noc);
end
150
function setup
% pass parameters from top level component to sub-components
Erev.T = c_T;
Erev.noc = c_noc;
% Ract.Area = c_Area;
% Ract.T = c_T;
Ract.noc = c_noc;
Rcon.Area = c_Area;
Rcon.B = c_B;
Rcon.Jmax = c_Jmax;
Rcon.noc = c_noc;
Rohm.Area = c_Area;
Rohm.T = c_T;
Rohm.l = c_l;
Rohm.noc = c_noc;
Cap.C = c_CapVal;
end
connections
connect(n, Erev.n, VSense.n);
connect(Erev.p, Ract.p, Cap.p);
connect(Ract.n, Rcon.p);
connect(Rcon.n, Rohm.p, Cap.n);
connect(Rohm.n, CSense.p);
connect(CSense.n, p, VSense.p);
connect(H2_pressure, Erev.H2_pressure, Ract.H2_pressure);
connect(Air_pressure, Erev.Air_pressure, Ract.Air_pressure);
connect(CSense.I, gFlow.I, sCurrent);
connect(VSense.V, sVoltage);
connect(gFlow.H2_flow, H2_flow);
connect(gFlow.Air_flow, Air_flow);
connect(Erev.Vrev, eVr);
connect(Rohm.effective_Res, eRo);
connect(Rcon.effective_Res, eRc);
connect(Ract.effective_Res, eRa);
connect(Cap.Vdl, Vdl);
connect(Cap.Idl, Idl);
end
end
151
APPENDIX M
152
% PEMFC_Dynamic_Test.m
%
% Test the PEMFC for a pulsed current load for a range of
% double layer capacitance values.
cdl = 10^-6;
cdl_vals = [10^-6 0.1 0.25 0.5 1];
cm = [0 0 102; 204 0 204; 0 255 0; 255 128 0; 255 0 0];
cm = (1/255)*cm;
for i = 1:5
cdl = cdl_vals(i);
sim('MyPEMFC_2_Dynamics');
tv = simlog.PEMFC_model_2.sVoltage.series.time;
vv = simlog.PEMFC_model_2.sVoltage.series.values;
ti = simlog.PEMFC_model_2.sCurrent.series.time;
ii = simlog.PEMFC_model_2.sCurrent.series.values;
subplot(1,2,1)
plot(tv,vv, 'LineWidth', 2, 'Color', [cm(i,1) cm(i,2) cm(i,3)]);
if i == 1
hold on;
end
subplot(1,2,2)
plot(ti, ii, 'LineWidth', 2, 'Color', [cm(i,1) cm(i,2) cm(i,3)]);
if i == 1
hold on;
end
end
subplot(1,2,1)
hold off;
title('PEMFC Stack Voltage versus Time');
xlabel('Time (seconds)');
ylabel('Voltage (V)');
legend('C = 1 uF', 'C = 0.1 F', 'C = 0.25 F', 'C = 0.5 F', 'C = 1 F');
set(gca, 'FontSize', 14);
grid;
subplot(1,2,2)
hold off;
title('PEMFC Current versus Time');
xlabel('Time (seconds)');
ylabel('Voltage (V)');
legend('C = 1 uF', 'C = 0.1 F', 'C = 0.25 F', 'C = 0.5 F', 'C = 1 F');
set(gca, 'FontSize', 14);
grid;
set(gcf, 'color', 'w');
set(findall(gcf, 'type', 'text'), 'FontSize', 18);
153
APPENDIX N
154
component RPMvsPower
% RPMvsPower
inputs
I = { 0, '1' }; % P:left
end
outputs
O = { 0, '1' }; % RPM:right
end
parameters(Size=variable)
xd = [0 25 50 75 100 125 150 175 200 225 250 275 300 325 350];
yd = [0 2168 2983 3597 4107 4552 4952 5316 5654 5969 6267 6549 6816 7073 7319];
end
equations
O == tablelookup(xd,yd,I,interpolation=cubic,extrapolation=nearest);
end
end
component TorquevsRPM
% TorquevsRPM
inputs
I = { 0, '1' }; % RPM:left
end
outputs
O = { 0, '1' }; % T:right
end
parameters(Size=variable)
xd = [0 2168 2983 3597 4107 4552 4952 5316 5654 5969 6267 6549 ...
6816 7073 7319];
yd = -(5.8/(2*3.14159))*[0 0.001835 0.002668 0.003318 0.003875 0.004370 ...
0.004821 0.005239 0.005630 0.005999 0.006349 0.006683 ...
0.007005 0.007313 0.007611];
end
equations
O == tablelookup(xd, yd, I, interpolation=cubic, ...
extrapolation=nearest);
end
end
component TorqueLoad
% Torque_Load
inputs
S = { 0, 'N*m'}; % S:bottom
end
variables(Access = protected)
w = { 0, 'rpm' }; % angular velocity (across variable)
t = { 0, 'N*m' }; % torque (through variable)
end
branches
t : C.t -> R.t; % through variable
end
equations
w == C.w - R.w;
t == S;
end
end
156
APPENDIX O
OBSERVER_1 DIAGRAM
157
Figure 60. Observer_1 Diagram
158
APPENDIX P
OBSERVER_2 DIAGRAM
159
Figure 61. Observer_2 Diagram
160
APPENDIX Q
161
Figure 62. Hybrid Power System Model Configuration Parameters
162
APPENDIX R
163
%
% flight_profile_03.m
%
% Flight Time: 6 hours
% (21,600 seconds + 200 seconds idle after landing)
TFT = 6*60*60 + 200; % total flight time
% Fuel Cell parameters
fcnoc = 20; % number of cells in the fuel cell
fcF = 96485; % Faraday's constant (coulombs/mole)
% Input gas pressures to fuel cell
h2p = 0.5; % hydrogen pressure (atm)
airp = 1.0; % air pressure (atm)
% Hydrogen fuel stored
h2Fuel = 400; % total hydrogen fuel available (Liters at STP)
% Battery parameters
BatInitSOC = 100; % initial battery state of charge
BatCapacity = 8; % Battery Capacity
% Battery charger
charger_on = 0; % turn decent charger on or off (0 = off, 1 = on)
charger_val = 0.4; % current setpoint for charger
% Hybrid parameter: zero = no battery assist during cruise
hybrid = 0.0; % hybrid level -> adjust CAOC setpoint
signalbuilder('PowerManagement_08/Flight_Profile','set', ...
'profile_01', 'group_01', throttle(:,1), throttle(:,2));
signalbuilder('PowerManagement_08/Balance_of_Plant/Fuel_Cell_BOP', ...
'set', 'bop_profile_01', 'group_01', ...
bop_res(:,1), bop_res(:,2));
165
APPENDIX S
166
% fp03_run1and2_plot_voltage_and_current.m
%
% Run PowerManagement_08.slx for hybrid parameters h = {0, 1.5} and
% plot the voltage and current responses on the same graphs.
% m-file dependency: flight_profile_03.m
%
cm = [0 0 0; 0 0 250; 250 0 0; 0 140 0]; % color matrix
cm = (1/255)*cm; % normalize color matrix
flight_profile_03; % run flight_profile_03.m (sets hybrid = 0)
sim('PowerManagement_08.slx', 'StopTime', num2str(21800)); % run simulation
% extract data
vfc_h0p0_t = simlog.PEMFC_model_2.sVoltage.series.time;
vfc_h0p0_v = simlog.PEMFC_model_2.sVoltage.series.values;
ifc_h0p0_t = simlog.PEMFC_model_2.sCurrent.series.time;
ifc_h0p0_v = simlog.PEMFC_model_2.sCurrent.series.values;
vbat_h0p0_t = simlog.LiPo_4S.BatteryVoltage.series.time;
vbat_h0p0_v = simlog.LiPo_4S.BatteryVoltage.series.values;
ibat_h0p0_t = simlog.LiPo_4S.BatteryCurrent.series.time;
ibat_h0p0_v = simlog.LiPo_4S.BatteryCurrent.series.values;
hybrid = 1.5; % change the hybrid parameter to 1.5
sim('PowerManagement_08.slx', 'StopTime', num2str(21800)); % run simulation
% extract data
vfc_h1p5_t = simlog.PEMFC_model_2.sVoltage.series.time;
vfc_h1p5_v = simlog.PEMFC_model_2.sVoltage.series.values;
ifc_h1p5_t = simlog.PEMFC_model_2.sCurrent.series.time;
ifc_h1p5_v = simlog.PEMFC_model_2.sCurrent.series.values;
vbat_h1p5_t = simlog.LiPo_4S.BatteryVoltage.series.time;
vbat_h1p5_v = simlog.LiPo_4S.BatteryVoltage.series.values;
ibat_h1p5_t = simlog.LiPo_4S.BatteryCurrent.series.time;
ibat_h1p5_v = simlog.LiPo_4S.BatteryCurrent.series.values;
% plot the results
fh = figure;
subplot(1,2,1)
plot(vfc_h0p0_t/60, vfc_h0p0_v, 'LineWidth', 2, 'Color', cm(1,:));
hold on;
plot(vbat_h0p0_t/60, vbat_h0p0_v, 'LineWidth', 2, 'Color', cm(2,:));
plot(vfc_h1p5_t/60, vfc_h1p5_v, 'LineWidth', 2, 'Color', cm(3,:));
plot(vbat_h1p5_t/60, vbat_h1p5_v, 'LineWidth', 2, 'Color', cm(4,:));
hold off;
title('Fuel Cell and Battery Voltage versus Time');
xlabel('Time (minutes)');
ylabel('Voltage (V)');
legend('Fuel Cell Voltage, hybrid = 0', 'Battery Voltage, hybrid = 0', ...
'Fuel Cell Voltage, hybrid = 1.5', 'Battery Voltage, hybrid = 1.5');
set(gca, 'FontSize', 14);
grid;
subplot(1,2,2)
plot(ifc_h0p0_t/60, ifc_h0p0_v, 'LineWidth', 2, 'Color', cm(1,:));
hold on;
plot(ibat_h0p0_t/60, ibat_h0p0_v, 'LineWidth', 2, 'Color', cm(2,:));
plot(ifc_h1p5_t/60, ifc_h1p5_v, 'LineWidth', 2, 'Color', cm(3,:));
plot(ibat_h1p5_t/60, ibat_h1p5_v, 'LineWidth', 2, 'Color', cm(4,:));
hold off;
title('Fuel Cell and Battery Current versus Time');
xlabel('Time (minutes)');
167
ylabel('Current (A)');
legend('Fuel Cell Current, hybrid = 0', 'Battery Current, hybrid = 0', ...
'Fuel Cell Current, hybrid = 1.5', 'Battery Current, hybrid = 1.5');
set(gca, 'FontSize', 14);
grid;
set(gcf, 'color', 'w');
set(findall(gcf, 'type', 'text'), 'FontSize', 18);
168
APPENDIX T
169
% fp03_run1and2_plot_Power_and_Wh.m
%
% Run PowerManagement_08.slx for hybrid parameters h = {0, 1.5} and
% plot the Fuel Cell, Battery, Load, Power, and Q1 Power on the same graph.
% m-file dependency: flight_profile_03.m
%
cm = [0 0 0; 0 0 250; 250 0 0; 0 140 0]; % color matrix
cm = (1/255)*cm; % normalize color matrix
flight_profile_03; % run flight_profile_03.m (sets hybrid = 0)
sim('PowerManagement_08.slx', 'StopTime', num2str(21800)); % run simulation
% extract data
% Fuel Cell and Battery Voltage and Current
vfc_h0p0_t = simlog.PEMFC_model_2.sVoltage.series.time;
vfc_h0p0_v = simlog.PEMFC_model_2.sVoltage.series.values;
ifc_h0p0_t = simlog.PEMFC_model_2.sCurrent.series.time;
ifc_h0p0_v = simlog.PEMFC_model_2.sCurrent.series.values;
vbat_h0p0_t = simlog.LiPo_4S.BatteryVoltage.series.time;
vbat_h0p0_v = simlog.LiPo_4S.BatteryVoltage.series.values;
ibat_h0p0_t = simlog.LiPo_4S.BatteryCurrent.series.time;
ibat_h0p0_v = simlog.LiPo_4S.BatteryCurrent.series.values;
FC_Power_h0p0 = vfc_h0p0_v .* ifc_h0p0_v; % calculate fuel cell power
Bat_Power_h0p0 = vbat_h0p0_v .* ibat_h0p0_v; % calculate battery power
% Elecrical load power and Mechanical load power
eLoad_h0p0_t = simlog.Motor_Prop_Subsystem.Electrical_Product.O.series.time;
eLoad_h0p0_v = simlog.Motor_Prop_Subsystem.Electrical_Product.O.series.values;
mLoad_h0p0_t = simlog.Motor_Prop_Subsystem.Mechanical_Product.O.series.time;
mLoad_h0p0_v = simlog.Motor_Prop_Subsystem.Mechanical_Product.O.series.values;
% Q1 Vds and current
q1vs_h0p0_t = simlog.Controller.Q1.S.v.series.time;
q1vs_h0p0_v = simlog.Controller.Q1.S.v.series.values;
q1vd_h0p0_t = simlog.Controller.Q1.D.v.series.time;
q1vd_h0p0_v = simlog.Controller.Q1.D.v.series.values;
q1vds_h0p0 = q1vs_h0p0_v - q1vd_h0p0_v; % calculate voltage drop of Q1
q1power_h0p0 = q1vds_h0p0 .* ibat_h0p0_v; % calculate Q1 power
hybrid = 1.5; % change the hybrid parameter to 1.5
sim('PowerManagement_08.slx', 'StopTime', num2str(21800)); % run simulation
% extract data
% Fuel Cell and Battery Voltage and Current
vfc_h1p5_t = simlog.PEMFC_model_2.sVoltage.series.time;
vfc_h1p5_v = simlog.PEMFC_model_2.sVoltage.series.values;
ifc_h1p5_t = simlog.PEMFC_model_2.sCurrent.series.time;
ifc_h1p5_v = simlog.PEMFC_model_2.sCurrent.series.values;
vbat_h1p5_t = simlog.LiPo_4S.BatteryVoltage.series.time;
vbat_h1p5_v = simlog.LiPo_4S.BatteryVoltage.series.values;
ibat_h1p5_t = simlog.LiPo_4S.BatteryCurrent.series.time;
ibat_h1p5_v = simlog.LiPo_4S.BatteryCurrent.series.values;
FC_Power_h1p5 = vfc_h1p5_v .* ifc_h1p5_v; % calculate fuel cell power
Bat_Power_h1p5 = vbat_h1p5_v .* ibat_h1p5_v; % calculate battery power
% Electrical load power and Mechanical load power
eLoad_h1p5_t = simlog.Motor_Prop_Subsystem.Electrical_Product.O.series.time;
eLoad_h1p5_v = simlog.Motor_Prop_Subsystem.Electrical_Product.O.series.values;
mLoad_h1p5_t = simlog.Motor_Prop_Subsystem.Mechanical_Product.O.series.time;
mLoad_h1p5_v = simlog.Motor_Prop_Subsystem.Mechanical_Product.O.series.values;
% Q1 Vds and current
q1vs_h1p5_t = simlog.Controller.Q1.S.v.series.time;
170
q1vs_h1p5_v = simlog.Controller.Q1.S.v.series.values;
q1vd_h1p5_t = simlog.Controller.Q1.D.v.series.time;
q1vd_h1p5_v = simlog.Controller.Q1.D.v.series.values;
q1vds_h1p5 = q1vs_h1p5_v - q1vd_h1p5_v; % calculate voltage drop of Q1
q1power_h1p5 = q1vds_h1p5 .* ibat_h1p5_v; % calculate Q1 power
% plot the results
fh = figure;
subplot(1,2,1)
plot(vfc_h0p0_t/60, FC_Power_h0p0, 'LineWidth', 2, 'Color', cm(1,:));
hold on;
plot(vbat_h0p0_t/60, Bat_Power_h0p0, 'LineWidth', 2, 'Color', cm(2,:));
plot(eLoad_h0p0_t/60, eLoad_h0p0_v, 'LineWidth', 2, 'Color', cm(3,:));
plot(ibat_h0p0_t/60, q1power_h0p0, 'LineWidth', 2, 'Color', cm(4,:));
hold off;
title('Electrical Power, hybrid = 0 (Total Energy = 583.5 Wh)');
xlabel('Time (minutes)');
ylabel('Power (W)');
legend('Fuel Cell Power', 'Battery Power', 'Load Power', 'Q1 Power');
set(gca, 'FontSize', 14);
grid;
subplot(1,2,2)
plot(vfc_h1p5_t/60, FC_Power_h1p5, 'LineWidth', 2, 'Color', cm(1,:));
hold on;
plot(vbat_h1p5_t/60, Bat_Power_h1p5, 'LineWidth', 2, 'Color', cm(2,:));
plot(eLoad_h1p5_t/60, eLoad_h1p5_v, 'LineWidth', 2, 'Color', cm(3,:));
plot(ibat_h1p5_t/60, q1power_h1p5, 'LineWidth', 2, 'Color', cm(4,:));
hold off;
title('Electrical Power, hybrid = 1.5 (Total Energy = 585.2 Wh)');
xlabel('Time (minutes)');
ylabel('Power (W)');
legend('Fuel Cell Power', 'Battery Power', 'Load Power', 'Q1 Power');
set(gca, 'FontSize', 14);
grid;
set(gcf, 'color', 'w');
set(findall(gcf, 'type', 'text'), 'FontSize', 18);
171
APPENDIX U
172
% fp03_run1and2_plot_BatSOC_and_Hydrogen.m
%
% Run PowerManagement_08.slx for hybrid parameters h = {0, 1.5} and
% plot the battery SOC and hydrogen consumed on the same graphs.
% m-file dependency: flight_profile_03.m
%
cm = [0 0 0; 0 0 250; 250 0 0; 0 140 0]; % color matrix
cm = (1/255)*cm; % normalize color matrix
flight_profile_03; % run flight_profile_03.m (sets hybrid = 0)
sim('PowerManagement_08.slx', 'StopTime', num2str(21800)); % run simulation
% extract data
soc_h0p0_t = simlog.LiPo_4S.SOC.series.time;
soc_h0p0_v = simlog.LiPo_4S.SOC.series.values;
hyd_h0p0_t = simlog.Observer_1.H2Liters.O.series.time;
hyd_h0p0_v = simlog.Observer_1.H2Liters.O.series.values;
hybrid = 1.5; % change the hybrid parameter to 1.5
sim('PowerManagement_08.slx', 'StopTime', num2str(21800)); % run simulation
% extract data
soc_h1p5_t = simlog.LiPo_4S.SOC.series.time;
soc_h1p5_v = simlog.LiPo_4S.SOC.series.values;
hyd_h1p5_t = simlog.Observer_1.H2Liters.O.series.time;
hyd_h1p5_v = simlog.Observer_1.H2Liters.O.series.values;
% plot the results
fh = figure;
subplot(1,2,1)
plot(soc_h0p0_t/60, soc_h0p0_v, 'LineWidth', 2, 'Color', cm(1,:));
hold on;
plot(soc_h1p5_t/60, soc_h1p5_v, 'LineWidth', 2, 'Color', cm(3,:));
hold off;
title('Battery SOC versus Time');
xlabel('Time (minutes)');
ylabel('SOC (%)');
legend('Battery SOC, hybrid = 0', 'Battery SOC, hybrid = 1.5');
set(gca, 'FontSize', 14);
grid;
subplot(1,2,2)
plot(hyd_h0p0_t/60, hyd_h0p0_v, 'LineWidth', 2, 'Color', cm(1,:));
hold on;
plot(hyd_h1p5_t/60, hyd_h1p5_v, 'LineWidth', 2, 'Color', cm(3,:));
hold off;
title('Hydrogen Consumed versus Time');
xlabel('Time (minutes)');
ylabel('Hydrogen (Liters)');
legend('Hydrogen Consumed, hybrid = 0', 'Hydrogen Consumed, hybrid = 1.5');
set(gca, 'FontSize', 14);
grid;
set(gcf, 'color', 'w');
set(findall(gcf, 'type', 'text'), 'FontSize', 18);
173
APPENDIX V
174
% fp03_run1and2_plot_Q1_Power.m
%
% Run PowerManagement_08.slx for hybrid parameters h = {0, 1.5} and
% plot the Q1 Power for h = {0, 1.5} on the same graph.
% Use zoom plot within plot to zoom into to area of interest [0 200 -2 8].
% m-file dependency: flight_profile_03.m
%
cm = [0 0 0; 0 0 250; 250 0 0; 0 140 0]; % color matrix
cm = (1/255)*cm; % normalize color matrix
flight_profile_03; % run flight_profile_03.m (sets hybrid = 0)
sim('PowerManagement_08.slx', 'StopTime', num2str(21800)); % run simulation
% extract data
% Battery Current
ibat_h0p0_t = simlog.LiPo_4S.BatteryCurrent.series.time;
ibat_h0p0_v = simlog.LiPo_4S.BatteryCurrent.series.values;
% Q1 Vds and current
q1vs_h0p0_t = simlog.Controller.Q1.S.v.series.time;
q1vs_h0p0_v = simlog.Controller.Q1.S.v.series.values;
q1vd_h0p0_t = simlog.Controller.Q1.D.v.series.time;
q1vd_h0p0_v = simlog.Controller.Q1.D.v.series.values;
q1vds_h0p0 = q1vs_h0p0_v - q1vd_h0p0_v; % calculate voltage drop of Q1
q1power_h0p0 = q1vds_h0p0 .* ibat_h0p0_v; % calculate Q1 power
hybrid = 1.5; % change the hybrid parameter to 1.5
sim('PowerManagement_08.slx', 'StopTime', num2str(21800)); % run simulation
% extract data
% Battery Current
ibat_h1p5_t = simlog.LiPo_4S.BatteryCurrent.series.time;
ibat_h1p5_v = simlog.LiPo_4S.BatteryCurrent.series.values;
% Q1 Vds and current
q1vs_h1p5_t = simlog.Controller.Q1.S.v.series.time;
q1vs_h1p5_v = simlog.Controller.Q1.S.v.series.values;
q1vd_h1p5_t = simlog.Controller.Q1.D.v.series.time;
q1vd_h1p5_v = simlog.Controller.Q1.D.v.series.values;
q1vds_h1p5 = q1vs_h1p5_v - q1vd_h1p5_v; % calculate voltage drop of Q1
q1power_h1p5 = q1vds_h1p5 .* ibat_h1p5_v; % calculate Q1 power
% plot the results
fh = figure;
plot(ibat_h0p0_t/60, q1power_h0p0, 'LineWidth', 2, 'Color', cm(1,:));
hold on;
plot(ibat_h1p5_t/60, q1power_h1p5, 'LineWidth', 2, 'Color', cm(3,:));
hold off;
title('Q1 Power versus Time');
xlabel('Time (minutes)');
ylabel('Q1 Power (W)');
axis([0 370 -10 110]);
legend('hybrid = 0', 'hybrid = 1.5');
set(gca, 'FontSize', 14);
grid;
% zoom plot within same plot
axes('position', [0.25 0.4 0.5 0.5]);
box on; % put a box around new pair of axes
plot(ibat_h0p0_t/60, q1power_h0p0, 'LineWidth', 2, 'Color', cm(1,:));
hold on;
plot(ibat_h1p5_t/60, q1power_h1p5, 'LineWidth', 2, 'Color', cm(3,:));
hold off;
175
axis([0 200 -2 8]);
set(gca, 'FontSize', 18);
grid;
set(gcf, 'color', 'w');
set(findall(gcf, 'type', 'text'), 'FontSize', 18);
176
APPENDIX W
177
% fp03_run1and2_Calculate_Total_Energy.m
%
% Run PowerManagement_08.slx for hybrid parameters h = {0, 1.5} and
% calculate the difference in energy produced by the sources, consumed
% by Q1 and the load during the flight profiles when h = {0, 1.5}.
% m-file dependency: flight_profile_03.m
%
cm = [0 0 0; 0 0 250; 250 0 0; 0 140 0]; % color matrix
cm = (1/255)*cm; % normalize color matrix
% h = 0 simulation run
flight_profile_03; % run flight_profile_03.m (sets hybrid = 0)
sim('PowerManagement_08.slx', 'StopTime', num2str(21800)); % run simulation
% extract data
% Fuel Cell Voltage
vfc_h0p0_t = simlog.PEMFC_model_2.sVoltage.series.time;
vfc_h0p0_v = simlog.PEMFC_model_2.sVoltage.series.values;
% Fuel Cell Current
ifc_h0p0_t = simlog.PEMFC_model_2.sCurrent.series.time;
ifc_h0p0_v = simlog.PEMFC_model_2.sCurrent.series.values;
% Battery Voltage
vbat_h0p0_t = simlog.LiPo_4S.BatteryVoltage.series.time;
vbat_h0p0_v = simlog.LiPo_4S.BatteryVoltage.series.values;
% Battery Current
ibat_h0p0_t = simlog.LiPo_4S.BatteryCurrent.series.time;
ibat_h0p0_v = simlog.LiPo_4S.BatteryCurrent.series.values;
% Q1 Vds and current
q1vs_h0p0_t = simlog.Controller.Q1.S.v.series.time;
q1vs_h0p0_v = simlog.Controller.Q1.S.v.series.values;
q1vd_h0p0_t = simlog.Controller.Q1.D.v.series.time;
q1vd_h0p0_v = simlog.Controller.Q1.D.v.series.values;
q1vds_h0p0 = q1vs_h0p0_v - q1vd_h0p0_v; % calculate voltage drop of Q1
q1power_h0p0 = q1vds_h0p0 .* ibat_h0p0_v; % calculate Q1 power
% get motor-propeller load power data
eLoad_h0p0_t = simlog.Motor_Prop_Subsystem.Electrical_Product.O.series.time;
eLoad_h0p0_v = simlog.Motor_Prop_Subsystem.Electrical_Product.O.series.values;
% get BOP load power data
vBOP_h0p0_t = simlog.Balance_of_Plant.BOP_Power.v1.series.time;
vBOP_h0p0_v = simlog.Balance_of_Plant.BOP_Power.v1.series.values;
iBOP_h0p0_t = simlog.Balance_of_Plant.BOP_Power.i1.series.time;
iBOP_h0p0_v = simlog.Balance_of_Plant.BOP_Power.i1.series.values;
% h = 1.5 simulation run
hybrid = 1.5; % change the hybrid parameter to 1.5
sim('PowerManagement_08.slx', 'StopTime', num2str(21800)); % run simulation
% extract data
% Fuel Cell Voltage
vfc_h1p5_t = simlog.PEMFC_model_2.sVoltage.series.time;
vfc_h1p5_v = simlog.PEMFC_model_2.sVoltage.series.values;
% Fuel Cell Current
ifc_h1p5_t = simlog.PEMFC_model_2.sCurrent.series.time;
ifc_h1p5_v = simlog.PEMFC_model_2.sCurrent.series.values;
% Battery Voltage
vbat_h1p5_t = simlog.LiPo_4S.BatteryVoltage.series.time;
vbat_h1p5_v = simlog.LiPo_4S.BatteryVoltage.series.values;
% Battery Current
ibat_h1p5_t = simlog.LiPo_4S.BatteryCurrent.series.time;
178
ibat_h1p5_v = simlog.LiPo_4S.BatteryCurrent.series.values;
% Q1 Vds and current
q1vs_h1p5_t = simlog.Controller.Q1.S.v.series.time;
q1vs_h1p5_v = simlog.Controller.Q1.S.v.series.values;
q1vd_h1p5_t = simlog.Controller.Q1.D.v.series.time;
q1vd_h1p5_v = simlog.Controller.Q1.D.v.series.values;
q1vds_h1p5 = q1vs_h1p5_v - q1vd_h1p5_v; % calculate voltage drop of Q1
q1power_h1p5 = q1vds_h1p5 .* ibat_h1p5_v; % calculate Q1 power
% get motor-propeller load data
eLoad_h1p5_t = simlog.Motor_Prop_Subsystem.Electrical_Product.O.series.time;
eLoad_h1p5_v = simlog.Motor_Prop_Subsystem.Electrical_Product.O.series.values;
% get BOP load power data
vBOP_h1p5_t = simlog.Balance_of_Plant.BOP_Power.v1.series.time;
vBOP_h1p5_v = simlog.Balance_of_Plant.BOP_Power.v1.series.values;
iBOP_h1p5_t = simlog.Balance_of_Plant.BOP_Power.i1.series.time;
iBOP_h1p5_v = simlog.Balance_of_Plant.BOP_Power.i1.series.values;
% Calculate the difference in total energy consumed by Q1 during flight
% profiles h = 0 and h = 1.5.
% note: h = 0 and h = 1.5 simulation runs have different number of data
% points. Data from each run must be integrated separately.
m = length(ibat_h0p0_t);
n = length(ibat_h1p5_t);
Q1_Energy_h0p0 = zeros(1,m);
Q1_Energy_Total_h0p0 = 0;
Q1_Energy_h1p5 = zeros(1,n);
Q1_Energy_Total_h1p5 = 0;
delta_t = 0;
delta_Q1_Energy = 0;
% Calculate total energy consumed by Q1 for h = 0 run.
for i = 2:m
delta_t = ibat_h0p0_t(i) - ibat_h0p0_t(i-1);
y = 0.5*(q1power_h0p0(i) + q1power_h0p0(i-1));
Q1_Energy_h0p0(i) = y * delta_t; % Joules
Q1_Energy_Total_h0p0 = Q1_Energy_Total_h0p0 + Q1_Energy_h0p0(i);
end
% Calculate total energy consumed by Q1 for h = 1.5 run.
for i = 2:n
delta_t = ibat_h1p5_t(i) - ibat_h1p5_t(i-1);
y = 0.5*(q1power_h1p5(i) + q1power_h1p5(i-1));
Q1_Energy_h1p5(i) = y * delta_t; % Joules
Q1_Energy_Total_h1p5 = Q1_Energy_Total_h1p5 + Q1_Energy_h1p5(i);
end
delta_Q1_Energy = Q1_Energy_Total_h1p5 - Q1_Energy_Total_h0p0;
delta_Q1_Energy = delta_Q1_Energy / 3600; % convert to Watt-hours
% display the difference in Q1 total energy
delta_Q1_Energy
% Next, calculate total energy consumed by the motor-propeller load for h = 0
m = length(eLoad_h0p0_t);
Load_Energy_h0p0 = zeros(1,m);
Load_Energy_Total_h0p0 = 0;
for i = 2:m
delta_t = eLoad_h0p0_t(i) - eLoad_h0p0_t(i-1);
y = 0.5*(eLoad_h0p0_v(i) + eLoad_h0p0_v(i-1));
Load_Energy_h0p0(i) = y * delta_t; % Joules
Load_Energy_Total_h0p0 = Load_Energy_Total_h0p0 + Load_Energy_h0p0(i);
179
end
% Now, calculate total energy consumed by the motor-propeller load for h = 1.5
n = length(eLoad_h1p5_t);
Load_Energy_h1p5 = zeros(1,n);
Load_Energy_Total_h1p5 = 0;
for i = 2:n
delta_t = eLoad_h1p5_t(i) - eLoad_h1p5_t(i-1);
y = 0.5*(eLoad_h1p5_v(i) + eLoad_h1p5_v(i-1));
Load_Energy_h1p5(i) = y * delta_t; % Joules
Load_Energy_Total_h1p5 = Load_Energy_Total_h1p5 + Load_Energy_h1p5(i);
end
delta_Load = Load_Energy_Total_h1p5 - Load_Energy_Total_h0p0;
delta_Load = delta_Load / 3600; % convert to Watt-hours
delta_Load
% Next, calculate the total energy consumed by the BOP load for h = 0
m = length(iBOP_h0p0_t);
BOP_Energy_h0p0 = zeros(1,m);
BOP_Energy_Total_h0p0 = 0;
for i = 2:m
delta_t = iBOP_h0p0_t(i) - iBOP_h0p0_t(i-1);
y = 0.5*(vBOP_h0p0_v(i) + vBOP_h0p0_v(i-1));
z = 0.5*(iBOP_h0p0_v(i) + iBOP_h0p0_v(i-1));
BOP_Energy_h0p0(i) = (y * z) * delta_t; % Joules
BOP_Energy_Total_h0p0 = BOP_Energy_Total_h0p0 + BOP_Energy_h0p0(i);
end
% Now, calculate the total energy consumed by the BOP load for h = 1.5
n = length(iBOP_h1p5_t);
BOP_Energy_h1p5 = zeros(1,n);
BOP_Energy_Total_h1p5 = 0;
for i = 2:n
delta_t = iBOP_h1p5_t(i) - iBOP_h1p5_t(i-1);
y = 0.5*(vBOP_h1p5_v(i) + vBOP_h1p5_v(i-1));
z = 0.5*(iBOP_h1p5_v(i) + iBOP_h1p5_v(i-1));
BOP_Energy_h1p5(i) = (y * z) * delta_t; % Joules
BOP_Energy_Total_h1p5 = BOP_Energy_Total_h1p5 + BOP_Energy_h1p5(i);
end
delta_BOP = BOP_Energy_Total_h1p5 - BOP_Energy_Total_h0p0;
delta_BOP = delta_BOP / 3600; % change to Watt-hours
delta_BOP
181
APPENDIX X
182
Motor: AXI 4120/14
Electronic Speed Controller: JETI Advance 70 Pro (5V < Vin < 15V)
Radio Receiver: Spektrum AR7000
Radio Transmitter: Spektrum DX7
Power Sources: DC power supply BK Precision 1670A, Motorcycle battery 12V
Figure 64. Oscilloscope Capture of ESC Voltage and Current Motor Signals (zoomed in)
183
Figure 65. Motor Measurements with DC Power Supply and No Propeller
184
Figure 66. Motor Measurements with DC Power Supply with Propeller
185
Figure 67. Motor Measurements with Motorcycle Battery with Propeller
186
APPENDIX Y
187
% HPMC_Battery_Capacity.m
%
% Run the Power_Management_08.slx model with flight_profile_03.
% The battery capacity is varied from 8 Ah to 12 Ah, and the
% amount of hydrogen consumed during the flight profile is recorded.
% Plot the battery SOC and hydrogen consumption curves for each run.
% file dependency: flight_profile_03.m
cm = [0 0 0; 0 0 250; 250 0 0; 0 140 0; 244 160 64]; % color matrix
cm = (1/255)*cm; % normalize color matrix
flight_profile_03; % generate Matlab variables for simulation
hybrid = 1.5; % set the hybrid parameter to 1.5
h2_used = zeros(1,5); % store the h2 used for each simulation
bat_min_SOC = zeros(1,5); % record minimum battery SOC for each simulation
for i = 1:5
BatCapacity = 8 + (i-1); % set the battery capacity
% run the simulation
sim('PowerManagement_08.slx', 'StopTime', num2str(21800));
% extract the data
% battery SOC
bSOC_t = simlog.LiPo_4S.SOC.series.time;
bSOC_v = simlog.LiPo_4S.SOC.series.values;
% hydrogen consumed
hc_t = simlog.Observer_1.H2Liters.O.series.time;
hc_v = simlog.Observer_1.H2Liters.O.series.values;
% total energy consumed
h2_used(i) = max(hc_v);
bat_min_SOC(i) = min(bSOC_v);
% create plot
subplot(1,2,1)
plot(bSOC_t/60, bSOC_v, 'LineWidth', 2, 'Color', cm(i,:));
if i == 1
hold on;
end
subplot(1,2,2)
plot(hc_t/60, hc_v, 'LineWidth', 2, 'Color', cm(i,:));
if i == 1
hold on;
end
end
% generate strings for legend
bat_8 = strcat('Battery: 8 Ah, H2: ', num2str(h2_used(1),'%.1f'));
bat_9 = strcat('Battery: 9 Ah, H2: ', num2str(h2_used(2),'%.1f'));
bat_10 = strcat('Battery: 10 Ah, H2: ', num2str(h2_used(3),'%.1f'));
bat_11 = strcat('Battery: 11 Ah, H2: ', num2str(h2_used(4),'%.1f'));
bat_12 = strcat('Battery: 12 Ah, H2: ', num2str(h2_used(5),'%.1f'));
subplot(1,2,1)
hold off;
title('Battery SOC versus Time');
xlabel('Time (minutes)');
ylabel('SOC (%)');
legend(bat_8, bat_9, bat_10, bat_11, bat_12);
set(gca, 'FontSize', 14);
grid;
subplot(1,2,2)
188
hold off;
title('Hydrogen Consumed versus Time');
xlabel('Time (minutes)');
ylabel('Hydrogen Consumed (L)');
legend(bat_8, bat_9, bat_10, bat_11, bat_12);
set(gca, 'FontSize', 14);
grid;
set(gcf, 'color', 'w');
set(findall(gcf, 'type', 'text'), 'FontSize', 18);
189
APPENDIX Z
190
% Energy_Analysis.m
%
% Integrate Load Electric Power and divide by time to obtain
% a graph of energy use per minute (Wh/min)
Motor_t = simlog.Motor_Prop_Subsystem.Electrical_Product.O.series.time;
Motor_v = simlog.Motor_Prop_Subsystem.Electrical_Product.O.series.values;
BOP_t = simlog.Balance_of_Plant.BOP_Power.v1.series.time;
BOP_voltage = simlog.Balance_of_Plant.BOP_Power.v1.series.values;
BOP_current = simlog.Balance_of_Plant.BOP_Power.i1.series.values;
BOP_v = BOP_voltage .* BOP_current;
Power_v = Motor_v + BOP_v;
% Plot the electrical load
plot(Motor_t/60, Power_v/60, 'LineWidth', 2); % (Wh/min)
title('Electrical Load Power versus Time');
xlabel('Time (minutes)');
ylabel('Power (Wh/min)');
set(gca, 'FontSize', 18);
grid;
set(gcf, 'color', 'w');
set(findall(gcf, 'type', 'text'), 'FontSize', 18);
% Calculate energy used during takeoff phase of flight profile
take_off_energy = 0;
m = find(Motor_t == 480); % get index where t = 480 seconds
for i = 2:m
delta_t = Motor_t(i) - Motor_t(i-1);
y = 0.5*(Power_v(i) + Power_v(i-1));
energy = y * delta_t; % Joules
Wh = energy / 3600;
take_off_energy = take_off_energy + Wh;
end
% Calculate energy used during landing phase of flight profile
landing_energy = 0;
n = find(Motor_t == 21200);
l = length(Motor_t);
for i = (n+1):l
delta_t = Motor_t(i) - Motor_t(i-1);
y = 0.5*(Power_v(i) + Power_v(i-1));
energy = y * delta_t; % Joules
Wh = energy / 3600;
landing_energy = landing_energy + Wh;
end
% display the results
[take_off_energy landing_energy]
191
APPENDIX AA
192
% Weight_FlightTime.m
%
% Compare flight time for various hybrid power system configurations.
% All configurations (except config_1) use all three sizes of Ninja
% compressed air tanks to store hydrogen (Ninja Pro v2 50, 68, & 90).
%
% config_1 = battery only
% config_2 = Hybrid with current assist, H-100 fuel cell, 6Ah battery
% config_3 = Hybrid with current assist, H-100 fuel cell, 11Ah battery
% config_4 = Hybrid with current assist, H-100 fuel cell, 17Ah battery
% config_5 = Hybrid with current assist, AST01-01 fuel cell, 6Ah battery
% config_6 = Hybrid with current assist, AST01-01 fuel cell, 11Ah battery
% config_7 = Hybrid with current assist, AST01-01 fuel cell, 17Ah battery
% config_8 = H-500 fuel cell only
% config_9 = Hybrid with no current assist, H-100 fuel cell, 6Ah battery
% using a 220g DC-to-DC converter (Conventional Hybrid)
%
% color matrix
cm = [0 0 0; 0 0 250; 250 0 0; 0 140 0; 244 160 64; 170 0 151];
cm = (1/255)*cm; % normalize color matrix
% Battery Data
Battery = [0.570 6; 0.850 11; 1.294 17]; % [kg Ah]
% Fuel Cell Data
FC_Weight = [0.55 1.69 2.92]; % [AST01-01 H-100 H-500] in kg
% DC-to-DC converter weight (400W => 0.220 kg)
DCC_Weight = 0.220; % (kg)
% Hydrogen Storage Tank Data
Ninja_50 = [0.936 166]; % [kg L]
Ninja_68 = [1.19 225.8]; % [kg L]
Ninja_90 = [1.45 298.9]; % [kg L]
% Second Stage Pressure Regulator
% Beswick engineering "Ultra-Miniature Single Stage Piston Regulator"
PR_Weight = 0.0062; % Pressure Regulator weight (Beswick)
% calculate config_2: Hybrid with current assist, H-100 fuel cell, 6Ah battery
wt_2 = zeros(1,6);
ft_2 = zeros(1,6);
wt_2(1) = Battery(1,1) + FC_Weight(2) + Ninja_50(1) + PR_Weight;
ft_2(1) = 18;
wt_2(2) = wt_2(1);
ft_2(2) = ((1.55*Ninja_50(2) + 10*Battery(1,2) - 44.9)/1.51) + 18;
wt_2(3) = Battery(1,1) + FC_Weight(2) + Ninja_68(1) + PR_Weight;
ft_2(3) = ft_2(2);
wt_2(4) = wt_2(3);
ft_2(4) = ((1.55*Ninja_68(2) + 10*Battery(1,2) - 44.9)/1.51) + 18;
wt_2(5) = Battery(1,1) + FC_Weight(2) + Ninja_90(1) + PR_Weight;
ft_2(5) = ft_2(4);
wt_2(6) = wt_2(5);
ft_2(6) = ((1.55*Ninja_90(2) + 10*Battery(1,2) - 44.9)/1.51) + 18;
% calculate config_3: Hybrid with current assist, H-100 fuel cell, 11Ah battery
wt_3 = zeros(1,6);
ft_3 = zeros(1,6);
wt_3(1) = Battery(2,1) + FC_Weight(2) + Ninja_50(1) + PR_Weight;
ft_3(1) = 18;
wt_3(2) = wt_3(1);
ft_3(2) = ((1.55*Ninja_50(2) + 10*Battery(2,2) - 44.9)/1.51) + 18;
wt_3(3) = Battery(2,1) + FC_Weight(2) + Ninja_68(1) + PR_Weight;
ft_3(3) = ft_3(2);
wt_3(4) = wt_3(3);
ft_3(4) = ((1.55*Ninja_68(2) + 10*Battery(2,2) - 44.9)/1.51) + 18;
wt_3(5) = Battery(2,1) + FC_Weight(2) + Ninja_90(1) + PR_Weight;
ft_3(5) = ft_3(4);
wt_3(6) = wt_3(5);
ft_3(6) = ((1.55*Ninja_90(2) + 10*Battery(2,2) - 44.9)/1.51) + 18;
% calculate config_4: Hybrid with current assist, H-100 fuel cell, 17Ah battery
wt_4 = zeros(1,6);
ft_4 = zeros(1,6);
wt_4(1) = Battery(3,1) + FC_Weight(2) + Ninja_50(1) + PR_Weight;
ft_4(1) = 18;
wt_4(2) = wt_4(1);
ft_4(2) = ((1.55*Ninja_50(2) + 10*Battery(3,2) - 44.9)/1.51) + 18;
wt_4(3) = Battery(3,1) + FC_Weight(2) + Ninja_68(1) + PR_Weight;
ft_4(3) = ft_4(2);
wt_4(4) = wt_4(3);
ft_4(4) = ((1.55*Ninja_68(2) + 10*Battery(3,2) - 44.9)/1.51) + 18;
wt_4(5) = Battery(3,1) + FC_Weight(2) + Ninja_90(1) + PR_Weight;
194
ft_4(5) = ft_4(4);
wt_4(6) = wt_4(5);
ft_4(6) = ((1.55*Ninja_90(2) + 10*Battery(3,2) - 44.9)/1.51) + 18;
% calculate config_5: Hybrid with current assist, AST01-01 fuel cell, 6Ah battery
wt_5 = zeros(1,6);
ft_5 = zeros(1,6);
wt_5(1) = Battery(1,1) + FC_Weight(1) + Ninja_50(1) + PR_Weight;
ft_5(1) = 18;
wt_5(2) = wt_5(1);
ft_5(2) = ((1.55*Ninja_50(2) + 10*Battery(1,2) - 44.9)/1.51) + 18;
wt_5(3) = Battery(1,1) + FC_Weight(1) + Ninja_68(1) + PR_Weight;
ft_5(3) = ft_5(2);
wt_5(4) = wt_5(3);
ft_5(4) = ((1.55*Ninja_68(2) + 10*Battery(1,2) - 44.9)/1.51) + 18;
wt_5(5) = Battery(1,1) + FC_Weight(1) + Ninja_90(1) + PR_Weight;
ft_5(5) = ft_5(4);
wt_5(6) = wt_5(5);
ft_5(6) = ((1.55*Ninja_90(2) + 10*Battery(1,2) - 44.9)/1.51) + 18;
% calculate config_6: Hybrid with current assist, AST01-01 fuel cell, 11Ah battery
wt_6 = zeros(1,6);
ft_6 = zeros(1,6);
wt_6(1) = Battery(2,1) + FC_Weight(1) + Ninja_50(1) + PR_Weight;
ft_6(1) = 18;
wt_6(2) = wt_6(1);
ft_6(2) = ((1.55*Ninja_50(2) + 10*Battery(2,2) - 44.9)/1.51) + 18;
wt_6(3) = Battery(2,1) + FC_Weight(1) + Ninja_68(1) + PR_Weight;
ft_6(3) = ft_6(2);
wt_6(4) = wt_6(3);
ft_6(4) = ((1.55*Ninja_68(2) + 10*Battery(2,2) - 44.9)/1.51) + 18;
wt_6(5) = Battery(2,1) + FC_Weight(1) + Ninja_90(1) + PR_Weight;
ft_6(5) = ft_6(4);
wt_6(6) = wt_6(5);
ft_6(6) = ((1.55*Ninja_90(2) + 10*Battery(2,2) - 44.9)/1.51) + 18;
% calculate config_7: Hybrid with current assist, AST01-01 fuel cell, 17Ah battery
wt_7 = zeros(1,6);
ft_7 = zeros(1,6);
wt_7(1) = Battery(3,1) + FC_Weight(1) + Ninja_50(1) + PR_Weight;
ft_7(1) = 18;
wt_7(2) = wt_7(1);
ft_7(2) = ((1.55*Ninja_50(2) + 10*Battery(3,2) - 44.9)/1.51) + 18;
wt_7(3) = Battery(3,1) + FC_Weight(1) + Ninja_68(1) + PR_Weight;
ft_7(3) = ft_7(2);
wt_7(4) = wt_7(3);
ft_7(4) = ((1.55*Ninja_68(2) + 10*Battery(3,2) - 44.9)/1.51) + 18;
wt_7(5) = Battery(3,1) + FC_Weight(1) + Ninja_90(1) + PR_Weight;
ft_7(5) = ft_7(4);
wt_7(6) = wt_7(5);
ft_7(6) = ((1.55*Ninja_90(2) + 10*Battery(3,2) - 44.9)/1.51) + 18;
% calculate config_9: Hybrid with no current assist, H-100 fuel cell, 6Ah battery
% using a 220g DC-to-DC converter (Conventional Hybrid)
wt_9 = zeros(1,6);
ft_9 = zeros(1,6);
wt_9(1) = Battery(1,1) + FC_Weight(2) + Ninja_50(1) + PR_Weight + DCC_Weight;
ft_9(1) = 18;
wt_9(2) = wt_9(1);
ft_9(2) = ((1.55*Ninja_50(2) - 44.9)/1.51) + 18;
wt_9(3) = Battery(1,1) + FC_Weight(2) + Ninja_68(1) + PR_Weight + DCC_Weight;
ft_9(3) = ft_9(2);
wt_9(4) = wt_9(3);
ft_9(4) = ((1.55*Ninja_68(2) - 44.9)/1.51) + 18;
wt_9(5) = Battery(1,1) + FC_Weight(2) + Ninja_90(1) + PR_Weight + DCC_Weight;
ft_9(5) = ft_9(4);
wt_9(6) = wt_9(5);
ft_9(6) = ((1.55*Ninja_90(2) - 44.9)/1.51) + 18;
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
% plot flight time versus weight
plot(ft_1, wt_1, 'LineWidth', 2, 'Marker', '*', 'Color', cm(1,:));
hold on;
plot(ft_2, wt_2, 'LineWidth', 2, 'Marker', '*', 'Color', cm(2,:));
plot(ft_3, wt_3, 'LineWidth', 2, 'Marker', '*', 'Color', cm(3,:));
plot(ft_4, wt_4, 'LineWidth', 2, 'Marker', '*', 'Color', cm(4,:));
ElevenAh = zeros(2,5);
wt_pt = Battery(2,1);
for i = 1:5
ElevenAh(1,i) = ((i*10*Battery(2,2) - 44.9)/1.51)+18;
ElevenAh(2,i) = i * wt_pt;
end
plot(ElevenAh(1,:), ElevenAh(2,:), 'o');
SeventeenAh = zeros(2,3);
wt_pt = Battery(3,1);
for i = 1:3
SeventeenAh(1,i) = ((i*10*Battery(3,2) - 44.9)/1.51)+18;
SeventeenAh(2,i) = i * wt_pt;
end
plot(SeventeenAh(1,:), SeventeenAh(2,:), 'o');
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
hold off;
title('Power System Weight versus Flight Time');
xlabel('Flight Time (minutes)');
ylabel('Power System Weight (kg)');
set(gca, 'FontSize', 18);
lh = legend('Battery Only', ...
'Hybrid: H-100, 6Ah Battery (Current Assist)', ...
'Hybrid: H-100, 11Ah Battery (Current Assist)', ...
'Hybrid: H-100, 17Ah Battery (Current Assist)', ...
'Hybrid: AST01-01, 6AhBattery (Current Assist)', ...
'Hybrid: AST01-01, 11AhBattery (Current Assist)', ...
'Hybrid: AST01-01, 17AhBattery (Current Assist)', ...
'H-500 Only', ...
'DC converter Hybrid: H-100, 6Ah Battery (No Current Assist)', ...
'Location', 'southeast');
grid;
set(gcf, 'color', 'w');
set(findall(gcf, 'type', 'text'), 'FontSize', 18);
set(lh, 'FontSize', 18);
197