Closed-Loop Thrust and Pressure Profile Throttling of A Nitrous-Oxide HTPB Hybrid Rocket Motor
Closed-Loop Thrust and Pressure Profile Throttling of A Nitrous-Oxide HTPB Hybrid Rocket Motor
Hybrid motors that employ non-toxic, non-explosive components with a liquid oxidizer
and a solid hydrocarbon fuel grain have inherently safe operating characteristics. The
inherent safety of hybrid rocket motors offers the potential to greatly reduce overall oper-
ating costs. Another key advantage of hybrid rocket motors is the potential for in flight
shutdown, restart, and throttle by controlling the pressure drop between the oxidizer tank
and the injector. The proposed research designs, develops, and ground tests a closed-loop
throttle controller for a hybrid rocket motor using nitrous oxide and hydroxyl-terminated
polybutadiene as propellants. The research simultaneously developed closed-loop throttle
algorithms and lab scale motor hardware to evaluate the fidelity of the throttle simulations
and algorithms. Initial open-loop motor tests were performed to better classify system
parameters and to validate motor performance values. Deep-throttle open-loop tests eval-
uated limits of stable thrust that can be achieved on the test hardware. Open loop tests
demonstrated the ability to throttle the motor to less than 10% of maximum thrust with
little reduction in effective specific impulse and acoustical stability. Following the open-
loop development, closed-loop, hardware-in-the-loop tests were performed. The closed loop
controller successfully tracked prescribed step and ramp and with a high degree of fidelity.
Nomenclature
1 of 42
Copyright © 2012 by Utah State University. Published by the American Institute of Aeronautics and Astronautics, Inc., with permission.
American Institute of Aeronautics and Astronautics
dp valve port diameter
e controller error
fa acoustical mode frequency
fl hybrid low frequency pole
Ft thrust
Ftf final thrust
Fti initial thrust
Ftn normalized thrust
fvs vortex shedding frequency
h1−3 enthalpies in plumbing feed physics model
Isp specific impulse
K total compressibility factor
KD derivative gain
Downloaded by "Indian Institute of Technology, Bombay" on March 31, 2025 | http://arc.aiaa.org | DOI: 10.2514/6.2012-4200
KI integral gain
KP proportional gain
l Strouhal number characteristic length
L∗ ratio of the chamber volume to the nozzle throat area
Lc combustion chamber length
Lf fuel grain length
ṁi injector mass flow rate
ṁp mass flow rate through valve port
ṁin mass flow rate through valve port in plumbing feed physics model
ṁout mass flow rate through injector in plumbing feed physics model
ṁox oxidizer mass flow rate
ṁtot total mass flow rate through motor
m2 post-throttle plumbing fluid mass in plumbing feed physics model
MB oxidizer bulk modulus
mAh milli-Amp hours
MVT maximum valve travel, see definition in Subsection 3.1.1
dn nozzle throat diameter
NHNE non-homogeneous, non-equilibrium
o/f oxidizer to fuel mass ratio
P̄ average pressure in plumbing
P̄c average chamber pressure
P̄i average injector manifold pressure
∆P pressure drop across orifice
P0 chamber pressure
P1 oxidizer feedline pressure in plumbing feed physics model
P2 post-throttle plumbing pressure in plumbing feed physics model
P0pc pre-combustion chamber pressure
RT product of gas constant and fluid temperature
RT̄ product of the average gas constant and temperature in the combustion chamber
RTc product of the gas constant and combustion temperature at the nozzle entrance
S controller thrust or chamber pressure signal
s differential operator
s1−3 entropies in plumbing feed physics model
Sr Strouhal number
t time
T1 oxidizer feedline temperature in plumbing feed physics model
T2 post-throttle plumbing temperature in plumbing feed physics model
ttot total controller run time
U Strouhal number fluid velocity
Uf fluid velocity in fuel grain
Uv fluid velocity through throttling valve
2 of 42
ρ density
ρ2 post-throttle plumbing fluid density in plumbing feed physics model
ρg average density of gas in plumbing
ρl average density of liquid in plumbing
ρox oxidizer density
τ time constant
τ1 delay transfer function time constant
τ2 delay transfer function time constant
τc combustion chamber fill time
τp plumbing fill time
I. Introduction
During the past 50 years conventional launch systems have been developed to a high state of capability;
however, for a variety of reasons these vehicles have become increasingly expensive to operate. Some of these
reasons include manufacturing and operational complexity, safety and environmental regulations for dealing
with hazardous materials, and the generally large “support army” required for flight preparations. Because
of high launch performance demands on specific impulse (Isp ) and thrust-to-weight ratio, conventional liquid
and solid-propelled rocket stages that employ highly-energetic, explosive, or toxic propellants will likely
remain the systems of choice for large military-class payloads or for human spaceflight. However, there exist
emerging markets, both commercial and government, that are willing to accept a lower system performance
in exchange for reduced operational costs and lower environmental impact.
3 of 42
with toxic propellant transport, storage, servicing, and clean up of accidental releases are becoming cost pro-
hibitive. As space flight operations continue to shift from government–run organizations to private companies
and universities that operate away from government-owned test reservations, servicing payloads requiring
hydrazine as a propellant becomes operationally infeasible. Extreme handling precautions generally do not
favor hydrazine as a propellant for secondary payloads.
A non-toxic, stable propellant alternative is clearly desired. Nitrous oxide (N2 O) is an inexpensive
and readily available propellant that has long been considered as a potential “green” monopropellant for
spacecraft applications. Despite a slight decrease in motor performance, N2 O has several advantages that
make it very competitive as a hybrid rocket oxidizer.
First, N2 O has the clear advantage of being non-toxic to human tissue and is classified as non-explosive
and non-flammable by the U.S. Occupational Safety and Health Administration (OHSA).7 Nitrous oxide
exists as a saturated liquid below its critical temperature of 36.4 C, and studies performed by the USAF
have demonstrated that it is virtually impossible to force a dissociation reaction with N2 O in its liquid form.8
Unlike the highly reactive propellants described in the previous paragraphs, nitrous oxide can be handled
without special precautions beyond those required for any pressurized fluid.
Second, unlike cryogenic liquid oxygen (LOX), N2 O is highly storable and allows rocket systems to
be prepared far in advance of motor use. Long-term propellant storability is a requirement for in-space
propulsion systems. Proposed hydrazine replacements based on aqueous solutions of hydroxyl-ammonium
nitrate (HAN) or hydroxyl-ammonium di-nitrate (ADN) salts are not long-term storable. The high water
content of the HAN/ADN propellants makes them susceptible to freezing during periods of extended cold
soak. These propellants must be temperature conditioned for space applications.
Finally, N2 O at room pressure exists in a saturated-liquid form and has a high vapor pressure exceeding
5000 kPa. This self-pressurizing property can be used to reduce complexity of propellant delivery systems. In
contrast hydrazine and peroxide have vapor pressures less than 1 kPa at room temperature. HAN/ADN and
LOX have essentially zero vapor pressure at storage temperature. It must be noted that the self-pressurizing
property of nitrous oxide typically produces two-phase injector flow, and this property has a significant
influence on the design of the throttling control system discussed in this paper. This effect will be described
in detail later.
While hybrid systems generally deliver lower Isp than conventional bipropellant liquid and lower volu-
metric efficiency than solid rockets of the same thrust level; because the propellant components remain inert
until ignited within the motor chamber, hybrid rockets are inherently safer to transport, load, store, and
operate. This inherent safety greatly reduces ground handling and transport costs, and can potentially lead
to overall reduction in system operating costs. Unlike solid-propelled rockets, where fuel grain flaws and
age-induced cracks present a significant safety issue, hybrid rockets exhibit a relative insusceptibility to grain
flaws.
4 of 42
While open-loop throttling can be accomplished with few additions to the oxidizer feed system, an open-
loop throttling system cannot adjust for motor variability. In addition to motor variability induced by
inconsistencies in manufacturing processes and operating conditions, hybrid rocket motors are inherently
variable by nature. Hybrid rocket thrust is more closely coupled with oxidizer mass flux than with chamber
pressure. Because mass flux is driven by surface friction along a turbulent boundary layer, hybrid motor
thrust is inherently chaotic. Some motor variability exists even if identical motors are used under identical
conditions. Closed-loop controlled throttling adds slightly more complexity to the throttling system than
does open-loop throttling, but allows the throttling system to compensate for motor variability. Controlled
throttling on a hybrid rocket allows for constant thrust or chamber pressure set points that greatly minimize
motor-to-motor thrust variability.
A high degree of motor-to-motor variability may be acceptable for experimental vehicles but will not
secure FAA certification for non-experimental, commercial spaceflight operations. Additionally, motor-to-
motor variability produces significant thrust asymmetries for clustered hybrid motor configurations, which
represents a significant hazard and currently precludes using multiple hybrid motors for launch vehicles.
Combustion physics require a longer aspect ratio (length to width ratio) for hybrid motors when compared
to their liquid and solid counterparts. Structural components associated with this high aspect ratio make
building a single hybrid motor that can produce sufficient thrust and impulse to achieve orbital velocity
difficult. For clustered motors, a closed-loop throttle controller can significantly reduce demand on launch
vehicle aerodynamic controls and structures.
5 of 42
gas N2 O system, 5) static-test demonstration of closed-loop thrust control on the N2 O-HTPB hybrid rocket
motor using sensed thrust feedback, and 6) static-test demonstration of closed-loop chamber pressure control
on the N2 O-HTPB hybrid rocket motor using sensed pre-combustion pressure transducer feedback.
Development of the closed-loop system simulation was necessary to provide an immediate prediction of
system behavior and eliminated extraneous system classification testing. The simulation allowed the system
gains to be quickly tuned by sweeping through a series of gains and observing the simulation response.
System parameters such as time constants of step-input responses, valve opening ranges, and instrument
noise were determined through open-loop testing. An open-loop deep throttle test was performed to evaluate
the limits of stable thrust in the test motors.
The first series of closed-loop throttle tests was performed on a nitrous oxide cold-gas system. Cold-flow
tests were performed first because of their reduced cost and complexity. Multiple tests could be performed
in succession without replacing fuel grains and ignitors, making it easier to compare results from multiple
gain sets to simulator results.
Following cold-flow testing, a thrust-feedback closed-loop throttle controller was demonstrated on a N2 O/
HTPB hybrid rocket motor. Much of the controller setup was the same as for the cold-flow tests. Only the
thrust profile used in the simulator and controller and the gains used for the controller needed to be altered.
Chamber pressure-feedback tests followed the thrust-feedback tests. The same basic form of controller was
used as for the thrust-feedback controllers with the appropriate terms modified for chamber pressure.
6 of 42
Control
Automation 85 Ω
Downloaded by "Indian Institute of Technology, Bombay" on March 31, 2025 | http://arc.aiaa.org | DOI: 10.2514/6.2012-4200
Linkage
0-5 V
Battery
Load Cell/
Control
Pressure Plumbing/Motor Physics
Valve
Transducer
7 of 42
effective port area provides a better representation of the linearity of a valve’s flow profile. Figure 2b shows
the effective port area as a function of valve position for a full-port ball valve. Although the flow profile of a
circular-port ball valve is non-linear, the open port provides the most weight efficient flow control ball valve.
Because demonstration of throttling control using a flight-weight valve was desired, a standard circular-port
ball valve was ultimately selected as the control valve.
100 1
30° V-port
90 60° V-port 0.9
90° V-port
80 0.8
Full Port
% Max. Effective Port Area
60 0.6
% Max. Cv
50 0.5
40 0.4
30 0.3
20 0.2
10 0.1
0 0
0 10 20 30 40 50 60 70 80 90 100 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
% MVT % MVT
(a) (b)
Figure 2: (a) Percent maximum Cv vs. opening percentage for different ball valve port geometries and (b)
Port area percentage vs. % MVT for a circular-port ball valve.
The valve was sized so that the pressure drop through the port was not sufficient to cause cavitation with
the valve fully open. This was mainly done to minimize flow losses and prevent unpredictable flow regimes
during non-throttled tests. During warmer months, the feedline pressure stays approximately 140 kPa above
the saturation pressure of nitrous oxide. To provide a safe margin, a pressure drop no more than 50 kPa
was allowed through the throttling valve when fully open. The port diameter required to maintain a given
pressure drop is given by38 s
4ṁp
dp = √ , (1)
πCdv 2ρ∆P
where ṁp is the mass flow rate through the port, Cdv is the discharge coefficient of the valve, ρ is the fluid
density, and ∆P is the pressure drop across the valve. Assuming a discharge coefficient of 0.8 for the valve,
a 7.8 mm orifice is required. The closest low-profile valve found was a 1/2” valve with a 5/16” (7.9 mm)
orifice sold by McMaster-Carra . The actual discharge coefficient was closer to 0.7 and the pressure drop was
around 70 kPa, but was still within the acceptable limit.
a McMaster-Carr Item# 45395K105, http://www.mcmaster.com/#catalog/118/433/=i4e9p2
8 of 42
signal. The 0-5 VDC analog servo input corresponds to a 0◦ -270◦ full range of servo rotation.
Control
Automation 85 Ω
0-20 mA 0-1.7 VDC Valve
Controller Resistor
Actuator
± 30 mV/
Linkage
0-5 V
Battery
Load Cell/
Control
Pressure Plumbing/Motor Physics
Valve
Transducer
For convenience, power is provided by a 12-volt lead acid battery for static testing. However, even at
peak draw the servo only requires 36 Watts of power. Controlling thrust for one minute of powered flight at
peak current draw would require a battery with rated to 50 mAh. It could be powered by a lithium polymer
battery that weighs only a fraction of a kilogram. Weighing in at about 1 kg, the valve and servo assembly
is potentially flight weight. Figure 3 shows the throttle valve assembly as it fits into the functional block
diagram in Fig. 1. Figure 4 shows the relation between the throttle valve assembly and the rest of the test
cart system piping.
b http://www.invenscience.com/index files/Page923.htm
c http://www.pololu.com/catalog/product/1392
9 of 42
4 Load Cell
5 7 Pressure Gauge
Pressure Transducer
Thermocouple
2 2 2 2 6 Burst Disk/ Safety Valve
Pneumatic Valve
Ball Valve
7
3
8 12
7
9
11 To Small Scale
Thrust Stand
1 1 1 10
1 N2O “K” Size Cylinder 7 Flex/Thin Tubing
2 He “K” Size Cylinder 8 Venturi Flow Meter
3 N2O Feed Manifold 9 98 mm Motor
4 He Feed Manifold 10 CO2 “K” Size Cylinder
2200 psi 800 psi 400 psi
5 He Regulator 11 Throttle Valve
6 Run Tank 12 Throttle Valve Actuator
To allow sufficient mass flow rates with minimal line losses, a predetermined mass of N2 O oxidizer,
nominally 500 g/s of burn time, was delivered to a closely coupled “run tank” from a series of “K” sized
industrial pressure cylinders. Nitrous oxide vapor pressure is highly dependent on temperature. Because the
test cell is open to the environment, the vapor pressure of the nitrous oxide in the MoNSTeR cart’s run tank
can be below saturation pressure if not controlled. Thus, the run tank was pressurized by gaseous helium
(He) to insure a constant feed pressure during the entire length of the burn. The He “top pressure” was set
by a manual regulator and was maintained near 5650 kPa for throttling tests.
The top pressure keeps the N2 O above saturation pressure for the entire run and insures a single-phase
liquid flow through the throttle valve for any expected ambient temperatures. The design motor chamber
pressure was 2760 kPa. A pneumatic run valve upstream of the throttle valve was triggered by an electronic
®
solenoid valve, and was automatically controlled by the instrumentation software.
Oxidizer mass flow was sensed by two vertical Omegadyne LCCD-100 (445 N) load cells mounted on
the run tank, and by an inline Venturi flow meter mounted in the oxidizer feed-line just ahead of the run
10 of 42
®
load cell and chamber pressure was sensed using an Omegadyne PX409-1.0KA5V (0-6900 kPa) absolute
pressure transducer mounted to the motor cap. An Omegadyne Type-K thermocouple was mounted at
the aft-end of the motor case to sense motor case temperature and thermal soak-back following the end of
the burn. All instrumentation was excited using a 10 VDC power source. The output response for the load
cells is 3 mV/Volt, and 0-5VDC for the pressure transducers.
The motors were mounted in a thrust balance on the MoNSTeR cart. A motor mounting bracket in the
thrust balance is supported on the sides by five ball and clevis joint linkages, two in the vertical and three
in the horizontal direction. Motion is constrained in the vertical and horizontal directions and rotations
are constrained about all three principal axes by the linkages. The axial load cell is attached between the
fore end of the motor mounting bracket and a rigid thrust beam using ball and clevis joints on either side.
Linkages and the axial load cell were aligned to within 0.2◦ of the principal axes using precision squares and
Downloaded by "Indian Institute of Technology, Bombay" on March 31, 2025 | http://arc.aiaa.org | DOI: 10.2514/6.2012-4200
inclinometers.
®
Two National Instruments data acquisition (DAQ) systems and control devices manage motor fire, control
throttling, and log test data. An NI-compact DAQ 9174d (cDAQ) 4-slot bus controller with multiple analog
input (16-bit), analog output, digital output, and thermocouple modules (24-bit) manage the majority of
the measurements and valve control. A National Instruments Compact Fieldpoint 2020e (cFP) automation
controller is used to manage the throttling valve controllers. An analog input module on the cFP allows
input measurements of thrust and chamber pressure for control loop feedback and to receive communication
signals from the cDAQ. Operators and experimenters are remotely located in a secure control room separated
from the test area. Communications to the test stand are managed by an operator-controlled computer via
®
universal serial bus (USB) using amplified extension cables. General control and measurement functions are
controlled by a LabVIEW virtual instrument (VI) hosted on the control computer. A separate throttle
valve controller VI was run on the cFP. Controller data was stored locally on the cFP and retrieved following
the test. Figure 5 presents a flowchart showing the communication signals between the instrumentation and
control devices.
The cFP automation controller was used to position the servo. The servo has a 270◦ total rotation range,
but only 90◦ was required to actuate the control valve. The 0-20 mA signal output of the 12-bit cFP analog
output module was converted to a 0-1.7 VDC input (5/3 of the 0-5 VDC command input for the 270◦ total
range) for slightly over 90◦ of control authority over the valve with a 0.022◦ resolution. The current to
voltage conversion of the analog output module control signal was accomplished by placing an 85 Ω resistor
in series with the signal wire. Position feedback of the servo to the cDAQ was provided by an internal rotary
d http://sine.ni.com/nips/cds/view/p/lang/en/nid/207535
e http://sine.ni.com/nips/cds/view/p/lang/en/nid/11572
11 of 42
D. Test Motor
For these tests a commercially available Cesaroni 98-mm solid-rocket motor was modified by replacing the
original motor cap with a custom-designed motor cap with a single port oxidizer injector. A threaded pressure
transducer port was also installed in the motor injector cap to allow for chamber pressure measurements.
The stock nozzle holder was replaced by a custom nozzle holder with a nozzle that had a larger throat
diameter than the stock nozzle holder could support. To reduce run-to-run variability due to nozzle erosion,
nozzles fabricated from a single piece of high-density graphite replaced the original manufacturer-supplied
phenolic nozzle.
The nozzle has a 4.2:1 expansion ratio and has a design throat diameter of 1.7 cm. Two Estes “mini
Downloaded by "Indian Institute of Technology, Bombay" on March 31, 2025 | http://arc.aiaa.org | DOI: 10.2514/6.2012-4200
A” class 10-gram solid rocket motors were inserted into the injector cap as ignitors. Electronic matches
burned by a 12 volt DC signal ignited these small motors. The ignitors were replaced after each test firing.
Additional advantages provided by this configuration are a ready-made flight-weight motor and the ability
to rapidly reload between motor tests.
®
®
HTPB fuel grains were cast using the commercially available Sartomer Poly bd R-45M polybutadiene
resin and PAPI 94 MDI curative. Sartomer R-45M has a polymerization factor of approximately 50 and
a molecular weight of 2800 kg/kg-mol.39 PAPI 94 is a polymethylene polyphenylisocyanate produced by
®
Dow Plastics Inc.40 The formulation contains methylene diphenylene diisocyanate (MDI) in proprietary
proportions. The curative has an average molecular weight of 290 kg/kg-mol. The nitrogen, carbon, oxygen
(N-C-O) bonds in the MDI react with the hydroxyl (OH) terminations in the polybutadiene resin to cure
the fuel grain.
For these tests carbon black was added to the mixture to insure opaqueness and prevent radiative heating
of the fuel grain and motor case liners. HTPB/MDI/carbon black mass proportions were set at 87%/ 12.5%/
0.5%, respectively. Past experience has determined that these proportions assure adequate fuel grain cure
and material hardness. The resin and curative were mixed in a commercial paint mixer that was sealed and
fitted so that the fuel mixture could be placed under a vacuum during the mixing process. A commercial
H-VAC vacuum pump was used to remove gas bubbles created in the fuel grain during the mixing process.
®
The de-gassed mixture was cast in cardboard sleeves with a 2.67 cm OD polyvinyl chloride (PVC) pipe used
as a mandrel. Before casting, the mandrel was coated with Ease Release 400 mold release agent to insure
proper release after the fuel grain cured.
Each fuel grain was approximately 57.15 cm in length, 8.26 cm in diameter, the initial fuel port diameter
is 2.67 cm, and post combustion chambers are 5.66 cm in diameter and 1.27 cm deep. The mean density
of the HTPB fuel grains used for these tests was approximately 966 kg/m3 , and the cast fuel grains had a
mean mass of 2.50 kg. Fig. 6 presents a schematic of the test motor.
12 of 42
Igniter
8.26 cm
5.66 cm
Port
2.67 cm
Igniter Port
Cap Fuel Grain Nozzle
Downloaded by "Indian Institute of Technology, Bombay" on March 31, 2025 | http://arc.aiaa.org | DOI: 10.2514/6.2012-4200
2.54 cm
Phenolic Grain
1.27 cm
Liner
2.54 cm
57.15 cm
1000
900 Motor
Shutoff
800
700
Dynamics
500 Transient
400
300
200
First Igniter
100
0
0 2 4 6 8 10 12 14
Time (s)
Figure 7: Thrust time histories for previous HTPB motor variability burns.
13 of 42
close in value to Popc , and Popc was consistently used for all controller parameters.
Non-linear Lowpass
Feedback
Cancellation Filter
Although the same control structure was used for all tests, three different operational gain ranges were
evaluated for the cold-flow controller and the hot-flow thrust-feedback and chamber pressure-feedback con-
trollers. Motor thrust and chamber pressure require a non-linear controller to be controlled directly due to
the non-linear relationship between the valve position and its effective port area. In order to simplify the
controller, non-linear cancellation was used to control the thrust or chamber pressure indirectly using the
linearly controlled valve position. The relationship between thrust or chamber pressure and valve position
was determined experimentally through open-loop testing. This relationship allows valve position to be
parametrized as a function of thrust or chamber pressure inputs. Thrust and chamber pressure command
and feedback signals were subsequently converted to an estimated valve position and the control loop was
closed around the error in valve position. By closing the loop around valve position error, the non-linearity
between the valve position and effective port area was separated from the controller and effectively canceled.
Large deviations in motor performance from the motors used during open-loop testing could potentially
cause instabilities in the controller through the non-linear cancellation, but variances in performance of the
test motor configuration were not large enough to cause instability.
In the controllers, especially the cold-flow controller, feedback noise was a significant factor in controller
stability. A lowpass filter was used to remove higher frequency noise from the feedback signal. Potential
high-frequency instabilities are discussed in Subsection IV.C. All controllers were run at a loop frequency of
100 Hz. The loop frequency was limited by the processing power of the cFP. Steps taken to reduce processing
load of the cFP included limiting data recording to when the control loop was running, building a data array
14 of 42
1. Cold-Flow Controller
Preliminary closed-loop throttle tests were performed using a top-pressured nitrous oxide cold-gas system.
Cold-flow tests were performed first because of their reduced cost and risk. Multiple tests could also be
performed in succession without replacing fuel grains and ignitors, making it easier to compare results from
multiple gain sets to simulator results. The cold gas system used for cold-flow testing consisted of all of the
same MoNSTeR Cart feedline piping used for hybrid motor testing and the motor cap and injector from the
98 mm test motor. Helium top-pressurized nitrous oxide was evacuated through the 3.8 mm orifice in the
injector to ambient conditions.
During testing the motor top cap was mounted to the MoNSTeR Cart thrust stand. The 2225 N axial
Downloaded by "Indian Institute of Technology, Bombay" on March 31, 2025 | http://arc.aiaa.org | DOI: 10.2514/6.2012-4200
load cell was intended to read much higher thrust and the noise level of the signal was a significant portion
of the maximum thrust that the cold gas system could produce. To minimize the effect of the signal noise
on controller performance, an integral controller was used. Only step input command profiles were used in
the cold-flow controller. Because of the high noise-to-signal ratio, ramp input response would be difficult to
discern.
A commanded thrust profile was pre-programmed and used as a tracking signal for the controller. To
synchronize the controller VI with the test computer VI, a thrust trigger was built into the controller. When
the controller VI was set to “active” mode in the test VI, the throttle valve was programmed to open to
an initial position and remain there while the thrust level was below the threshold. At the end of the test
the same threshold value was used to signal the controller to stop. A heavily filtered thrust signal was used
to avoid controller initiation due to noise spikes. Thrust levels were low enough during cold-flow testing
that controller overshoot could cause the thrust level to dip below the threshold and turn off the controller
mid-test. To avoid terminating the controller mid-test, the thrust trigger was set just above the normal noise
level and the lower command setting of the controller was high enough to leave a buffer zone. The highest
noise level before test initiation was around 5 N, so the thrust trigger was set to 10 N. Lower command
thrust levels were set no lower than 30 N to ensure that the threshold value was not reached during the test
due to controller overshoot.
Logarithmic functions of the form
Ft
x = − ln − + 1 τ + x0 (2)
A
were used in the non-linear cancellation, where x is the valve position in %MVT, Ft is the thrust, and the
other variables are experimentally derived curve fit coefficients. Equation (2) is the inverse of Eq. (22). To
ensure that values returned from the functions were real and finite, an upper thrust limit was set to a value
just smaller than A. If the measured thrust exceeds the limit, the signal is reset to the limit value.
15 of 42
demonstrate the controller and required fewer changes to the controller VI. The same thrust trigger was
used to initiate the controller as in the thrust-feedback controller VI. The simulator also required slight
modification to model the chamber pressure-feedback controller.
1. Servomotor
Delays in the servo response were due to the maximum servo speed and the settling time of the built-in
positioning controller. As provided, the factory preset control gains of the internal PID servo position
controller produced a sluggish response that was insufficient for throttling. To speed up valve actuation
times, the gains were modified so that the internal position controller was critically damped. Figure 9
compares the servo response time for a 10% MVT change in position using the original and the modified
gains. Note that the servo response with the original gains was measured using a lower resolution external
potentiometer so the apparent steps in the response are measurement artifacts and do not reflect the actual
response.
With the modified gains, the time constant for valve repositioning was 0.2 seconds and maximum travel
speed was 60% MVT per second. For the range of valve positions required for this project, the maximum
delay due to the servo was estimated to be 0.8 seconds.
16 of 42
10
8
% MVT
6
Downloaded by "Indian Institute of Technology, Bombay" on March 31, 2025 | http://arc.aiaa.org | DOI: 10.2514/6.2012-4200
2
Original Gains
Modified Gains
0
0 0.2 0.4 0.6 0.8 1 1.2
Time (s)
where Xox is the quality of the oxidizer, γox is the oxidizer ratio of specific heats, ρl is the liquid phase
density of the oxidizer, and P̄ is the average pressure in the system piping. Because a single port injector
was used for this project, there was no injector manifold and the same values were used for P̄ and P̄i . The
liquid compressibility factor, βs , is the inverse of the bulk modulus of the oxidizer, defined as42
P2 − P1
MB = , (6)
1 − ρ1 /ρ2
which describes the difference in the amount that a substance compresses between two different pressures, P1
and P2 ; ρ1 and ρ2 are the densities corresponding to those pressures. The bulk modulus has a nearly linear
relationship with temperature. Using temperatures and pressures in the ranges expected in the oxidizer feed
line, a linear function of temperature was developed to approximate MB .
Two-phase thermodynamic properties were determined using a Helmholtz model.43–45 Measured temper-
atures, pressures, and mass flow rates were assumed to be representative of the average values in the system
piping. The calculated values of τp ranged between 0.02 and 0.14 seconds for the full throttling range.
17 of 42
fc (γ) RTc L∗
τc = , (7)
Cdn RT̄ c∗
where Cdn is the nozzle discharge coefficient, RTc is the product of the gas constant and combustion tem-
perature at the nozzle entrance, RT̄ is the product of the average gas constant and temperature in the
combustion chamber, L∗ is the ratio of the chamber volume to the nozzle throat area, c∗ is the characteristic
velocity, and
(γ+1)/(γ−1)
ηc γ + 1
fc (γ) = (8)
γ 2
Downloaded by "Indian Institute of Technology, Bombay" on March 31, 2025 | http://arc.aiaa.org | DOI: 10.2514/6.2012-4200
where ηc is the combustion efficiency and γ is the ratio of specific heats of the combustion products. Com-
bustion properties c∗ , Tc , T̄ , γ, and the average and nozzle molecular weights were calculated using NASA’s
Chemical Equilibrium with Applications (CEA) program.46, 47 The nozzle discharge coefficient was esti-
mated to be 0.98.38 Previous testing showed the combustion efficiency to be about 0.97 for this motor
configuration.33 The calculated combustion chamber fill time was 0.01 seconds.
7. Vortex Shedding
The Strouhal number is defined as48
fvs l
Sr = , (11)
U
where fvs is a vortex shedding frequency, l is the characteristic length, and U is the fluid velocity. The
vortex shedding frequency is calculated by rearranging Eq. (11) to solve for fvs . In hybrid rocket motors,
18 of 42
where ṁtot is the total mass flow rate of fluid at the fuel grain exit. The product of the gas constant and
temperature at the fuel grain exit was assumed to be approximately equal to RT̄ and l was assumed to be
Downloaded by "Indian Institute of Technology, Bombay" on March 31, 2025 | http://arc.aiaa.org | DOI: 10.2514/6.2012-4200
the exit diameter of the fuel grain. Based on these parameters Eq. (13) calculates that the lowest vortex
shedding frequency at the fuel grain exit occurs at approximately 350 Hz.
9
LC noise
8 Test Stand Impulse
Resonance
7
6
Y(Thrust) N
4 Unidentified
Noise
3
0
0 50 100 150 200 250 300 350 400 450 500
Frequency (Hz)
19 of 42
1 2
20 of 42
3
1 In1
DerGainOut
Gain Derivative
2nd-Order
Filter3 1
K Ts (z+1)
Ki Out1
2(z-1)
Gain3 Discrete-Time
Integrator IntegralGainOut
To Workspace3
Kp
Gain4 PropOut
To Workspace
Figure 12: Simulink PID controller.
The PID controller gains were originally modeled based on the LabVIEW PID controller block to pro-
vide better agreement between the simulator and the controller VI.50 Later, the LabVIEW PID block was
abandoned to provide greater flexibility and the gains were converted to standard format, but the LabVIEW
signal flow path remained. Figure 13 shows how the gain model was integrated into the simulator control
loop.
Command 𝐾𝑃 + System
+
Profile − Dynamics
𝐾𝐼 Trapezoidal
Integrator
Non-linear
+
Cancellation −
𝑑
𝐾𝐷
𝑑𝑡
Non-linear Lowpass
Feedback
Cancellation Filter
21 of 42
where τ1 and τ2 are time constants acquired during open-loop testing as outlined in Subsection VI.C below.
The maximum servo speed was modeled by saturating the rate of change in the servo position. Figure 14
shows the simulator servo model.
Downloaded by "Indian Institute of Technology, Bombay" on March 31, 2025 | http://arc.aiaa.org | DOI: 10.2514/6.2012-4200
ServoFeedback 1
Out2
To Workspace7
1
1 Gservoloop du/dt
s
In1
LTI System Derivative1 Saturation2 Integrator1 Saturation3
and dividing it by the internal volume of the post-throttle system piping, V2 . Temperature was determined
using the Newton solver to match the enthalpies of State 1 and State 2. The pressure in the post-throttle
system piping is calculated as the saturation pressure with the density in the system piping and at the
temperature at which the isenthalpic condition is satisfied.
22 of 42
𝐶𝑑𝑣 𝐴𝑒𝑣 𝑃1 𝑇1 𝑋1
Injector Feedline
Model
Non-HEM Valve
Port Model
𝑃2
+
−
Integrator
Non-HEM Injector
Orifice Model
𝑚2
𝑚𝑜𝑢𝑡
𝑃1 𝑃2
ℎ1 ℎ2 𝐶𝑑𝑣 ℎ3 𝐶𝑑𝑖𝑛𝑗
𝑇2 𝑚2 ℎ1 ℎ1
𝑇1 𝑠1 𝐴𝑒𝑣 𝑠2 𝐴𝑖𝑛𝑗 𝑃0
𝑠1 𝑋2 𝜌2 𝑠2 𝑠3
𝑚𝑖𝑛 𝑚𝑜𝑢𝑡
𝑋1 𝑉2
Figure 16: Thermodynamic properties used in system piping feed physics model.
Conditions computed by the first NHNE model and the post-throttle system piping pressure were used
in a second NHNE model to compute the mass flow rate through the injector (ṁin ) as a function of P2 , s2 ,
and P0 . Chamber pressure has an approximately linear relationship with injector mass flow rate in the test
motor configuration. Experimental data was used to produce the equation,
P0 = a2 ṁout + b2 , (16)
that calculates chamber pressure based on ṁout , where a2 and b2 are experimentally derived curve fit
coefficients described further in Subsection VI.D. For the pressure-feedback controller model, this chamber
pressure value was output to the simulator.
23 of 42
For the thrust-feedback controller, another experimentally derived equation described in Section VI.D
was used to calculate thrust based on the chamber pressure. Thrust had a slightly quadratic relationship
with chamber pressure in the test motor configuration. The equation used to compute thrust is
Ft = a1 P02 + b1 P0 + c1 . (17)
When the full system piping model was placed in the simulator run times became excessive. In order to
decrease simulator run times, data from the full system piping model was used to make lookup table blocks
so that the imbedded MATLAB functions could be bypassed. The lookup table version of the system piping
model is shown in Fig. 17.
AValveOut1
1 To Workspace8
ValveArea
mdotoxOut1
Product To Workspace1
f(u)
1
P22Gtot
s
m0 xo
u2
2-D Lookup
Table mdotinjOut1
1 To Workspace
f(u)
tauMotor.s+1
Transfer Fcn Fcn
1
ChamberPressure pPlumbOut1
To Workspace2
f(u) -K-
rhoP2PP 1/Vol
24 of 42
Bandpass
Band-Limited
Downloaded by "Indian Institute of Technology, Bombay" on March 31, 2025 | http://arc.aiaa.org | DOI: 10.2514/6.2012-4200
Bandpass Filter
White Noise1
1
Bandpass
Out1
Band-Limited
Bandpass Filter1
White Noise2
Bandpass
Band-Limited
Bandpass Filter2
White Noise3
Band-Limited
White Noise4
Figure 18: Pressure transducer noise model.
5
Simulated Noise
4.5 Measured Noise
3.5
PSD (kPa2/Hz)
2.5
1.5
0.5
0
0 50 100 150 200 250 300 350 400 450 500
Frequency (Hz)
25 of 42
1 2 3 4
3000
Downloaded by "Indian Institute of Technology, Bombay" on March 31, 2025 | http://arc.aiaa.org | DOI: 10.2514/6.2012-4200
2000
1
1000
0
3000
Chamber Pressure (kPa)
2000
2 1000
0
3000
2000
3 1000
0
3000
2000
4
1000
0
0 5 10 0 5 10 0 5 10 0 5 10
Time (s)
Row\Col. 1 2 3 4
1 KP = 0.2, KI = 2.0 KP = 0.2, KI = 2.5 KP = 0.2, KI = 3.0 KP = 0.2, KI = 3.5
2 KP = 0.3, KI = 2.0 KP = 0.3, KI = 2.5 KP = 0.3, KI = 3.0 KP = 0.3, KI = 3.5
3 KP = 0.4, KI = 2.0 KP = 0.4, KI = 2.5 KP = 0.4, KI = 3.0 KP = 0.4, KI = 3.5
4 KP = 0.5, KI = 2.0 KP = 0.5, KI = 2.5 KP = 0.5, KI = 3.0 KP = 0.5, KI = 3.5
When a set of gains was selected it was checked to verify that it would work if motor performance varied
from the moderate performance parameters that were used for non-linear cancellation in the controller and
for the system piping model in the simulator. Motor variability was modeled by applying a scaling factor to
the effective valve port area that was input into the system piping model. The scaling factor was applied as
a gain following the table lookup for valve effective area, shown in Fig. 11. Mean motor thrust and chamber
pressure 2-σ variance was less than 8% for all tests performed to date on the 98 mm motor configuration. The
variability was modeled by scaling to ±15% of the effective valve port area, which creates a ±8% variance in
the thrust and chamber pressure. Figure 21 shows nominal simulator results compared to results from 85%
and 115% scaling. Comparison of Figures 20 and 21 shows that motor variability has a similar effect on the
controller to varying the integral gain.
26 of 42
2500
1500
1000
500
0
0 2 4 6 8 10 12
Time (s)
Downloaded by "Indian Institute of Technology, Bombay" on March 31, 2025 | http://arc.aiaa.org | DOI: 10.2514/6.2012-4200
3000
2500
Chamber Pressure (kPa)
2000
1500
1000
500
0
0 2 4 6 8 10 12
Time (s)
3000
2500
Chamber Pressure (kPa)
2000
1500
1000
500
0
0 2 4 6 8 10 12
Time (s)
Figure 21: Controller response for three different scaled valve areas.
27 of 42
80
70 90%
70% 80%
60%
60
50 50%
Downloaded by "Indian Institute of Technology, Bombay" on March 31, 2025 | http://arc.aiaa.org | DOI: 10.2514/6.2012-4200
Thrust (N)
40
30
40% Dsd
20
10
0 30%
0% 10% 20%
-10
0 2 4 6 8 10 12 14 16 18 20
Time (s)
28 of 42
were ignored and average values τ1 = 0.0221 s and τ2 = 0.0217 s were used for both cold-flow and hot-flow
simulation. Time constants were the same for thrust and chamber pressure, so the same values were used in
both simulator models. The natural frequency was calculated as part of the curve fit, but was not used in
the simulator.
1000 3500
900
3000
800
700 2500
600
2000
Thrust (N)
P0 (kPa)
500
1500
400
300 1000
200
100% MVT 500 100% MVT
100 50-90% MVT 50-90% MVT
25-65% MVT 25-65% MVT
0 0
0 2 4 6 8 10 12 14 0 2 4 6 8 10 12 14
Time (s) Time (s)
(a) (b)
Figure 23: (a) Thrust and (b) chamber pressure during hot flow open loop throttle tests.
1.2 1.2
1 1
0.8 0.8
Normalized Thrust
Normalized Thrust
0.6 0.6
0.4 0.4
0.2 0.2
0 curve 1 0 curve 1
curve 2 curve 2
curve 3 curve 3
-0.2 -0.2
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
t (sec) t (sec)
(a) (b)
Figure 24: Normalized thrust response curves for (a) cold-flow throttling and (b) hot-flow throttling.
29 of 42
Ft = a1 P02 + b1 P0 + c1 (20)
was used to describe the relationship between thrust (N) and absolute chamber pressure (kPa). The rela-
tionship between chamber pressure and oxidizer mass flow rate is mostly linear. A linear curve fit of the
form
P0 = a2 ṁox + b2 (21)
Downloaded by "Indian Institute of Technology, Bombay" on March 31, 2025 | http://arc.aiaa.org | DOI: 10.2514/6.2012-4200
was used to describe the relationship between absolute chamber pressure (kPa) and mass flow rate (kg/s).
Thrust can be related to mass flow rate by combining both relationships. Table 4 presents the values of
the curve fit coefficients used in Equations (20) and (21). Figure 25 shows the curve fit data compared to
experimental data. The experimental data shown in Fig. 25a was collected from all of the hot-fire open-loop
tests performed. While o/f shift occurred during each motor burn, it was not a significant factor in the
relationship between thrust and chamber pressure.
Table 4: Curve fit coefficients for thrust, chamber pressure, and mass flow rate calibrations.
Coefficient Value
a1 9.21 × 10−6
b1 0.28
c1 -40.15
a2 8602.47
b2 152.02
were used for the curve fits, where S is the controller input signal (thrust or absolute chamber pressure), x
is the valve position in %MVT, and the other variables are exponential curve fit coefficients that represent
scaling (A), initial offset (x0 ), and time constant (τ ). Equation 22 was rearranged into Eq. (2) for use in
the controllers. Values of the curve fit coefficients used for each controller are displayed in Table 5.
30 of 42
Thrust (N)
600
400
200
0
0 500 1000 1500 2000 2500 3000
Chamber Pressure (kPa)
(a)
Downloaded by "Indian Institute of Technology, Bombay" on March 31, 2025 | http://arc.aiaa.org | DOI: 10.2514/6.2012-4200
4000
Chamber Pressure (kPa)
Test Data
3000 Curve Fit
2000
1000
0
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35
Mass Flow Rate (kg/s)
(b)
Figure 25: Chamber pressure, thrust and mass flow calibrations compared to experimentally measured values.
800
600
Thrust (N)
400
3000
Chamber Pressure (kPa)
2000
1000
Curve Fit
Experimental Data
0
0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
% MVT
(b)
Figure 26: Experimentally derived curve fits for (a) thrust and (b) chamber pressure as a function of %
MVT.
31 of 42
Controller A x0 τ
Cold-Flow 75.04 -0.2527 0.2116
Thrust 819.1 -0.2268 0.1586
Chamber Pressure 2781 -0.2185 0.1522
Normalized Voltage/Servo Position
1.02 Voltage
Position
1.01
Downloaded by "Indian Institute of Technology, Bombay" on March 31, 2025 | http://arc.aiaa.org | DOI: 10.2514/6.2012-4200
1
0.99 ĞĂĚďĂŶĚ
0.98
0.97
0.96
-4 -3 -2 -1 0 1 2 3 4 5
Time (s)
32 of 42
1000
Raw Thrust
900 (1) Filtered Thrust
800
(4)
700
600
Thrust (N)
500
400
300
200
(2)
(5)
100 (3) (5)
(4)
0
0 5 10 15 20 25 30
Time (s)
Figure 28: Thrust time history with video stills from deep throttle test.
The trade-off for the ease and depth of hybrid rocket motor throttling is the loss of efficiency. Motor Isp
decreased significantly below 10% of peak thrust. Several techniques have been developed to counteract this
drop in efficiency such as oxidizer injection near the nozzle and pressure sensitive fuels.15 However, because
this project was considered an initial evaluation and the main focus was on developing a closed-loop throttle
controller, no steps were taken to improve deep throttling efficiency.
33 of 42
/ŶƐĞƚŽĨ&ŝŐ͘Ϯϴ
0.35
Raw Thrust
0.3 Filtered Thrust
Raw P0
0.25
Filtered P0
0.2
Normalized Thrust / P0
Downloaded by "Indian Institute of Technology, Bombay" on March 31, 2025 | http://arc.aiaa.org | DOI: 10.2514/6.2012-4200
0.15
0.1
0.05
-0.05
KdžŝĚŝnjĞƌ
^ƚĂďůĞ hŶƐƚĂďůĞ
-0.1 sĂůǀĞ
ŽŵďƵƐƚŝŽŶ ŽŵďƵƐƚŝŽŶ
ůŽƐĞĚ
-0.15
23 24 25 26 27 28 29
Time (s)
Figure 29: Normalized thrust and chamber pressure during unstable combustion in deep throttle test.
3.5
2.5
Injector Pressure Ratio
1.5
0.5
liquid rocket lower limit
0
0 5 10 15 20 25 30 35
Time (s)
34 of 42
The cutoff frequency, fc , is the value used in the lowpass filter for each controller. For thrust-feedback
tests (cold-flow and hot-flow) the cutoff frequency needed to at least be below 30 Hz to avoid having the test
stand structural harmonic frequency feed through the load cell into the controller. In both the cold-flow and
hot-flow tests, however, the cutoff frequency was set much lower to reduce the noise amplitude. The hybrid
low frequency mode was not observed in the test motor configuration, so resonance was not as much of a
concern for chamber pressure-feedback tests as it was for thrust-feedback tests. As with the thrust-feedback
tests, the cutoff frequency was selected mainly to reduce noise amplitude.
1. Thrust-Feedback Tests
Figures 32 and 33 show the results from Tests HTPBCL1 and HTPBCL2. HTPBCL1 was a step input
test with a 10 second duration. HTPBCL2 had a ramp input and a 15 second duration. Cold-flow closed-
loop testing provided a reasonable analog to the hot-flow thrust-feedback tests because the same feedback
mechanism was used. Both controllers performed as expected based on simulator and cold-flow test results.
The step profile tests had ringing in the controller response on the second step as Fig. 32 shows. On
the first step the servo was initialized to a position near where it settled during the initial set point. On
the second step the set point transitioned from a control region where thrust changes nearly linearly with
valve position to a region where large changes in valve position produce little change in thrust. The ringing
may be caused by an increasing controller signal during the transition. Changes in thrust from the nominal
35 of 42
100
80
Thrust (N) 60
40
Raw Thrust
20
Downloaded by "Indian Institute of Technology, Bombay" on March 31, 2025 | http://arc.aiaa.org | DOI: 10.2514/6.2012-4200
Filtered Thrust
Command
0
0 2 4 6 8 10 12 14
Time (s)
(a)
100
80
60
Thrust (N)
40
Raw Thrust
20
Filtered Thrust
Command
0
0 2 4 6 8 10 12
Time (s)
(b)
100
80
60
Thrust (N)
40
Raw Thrust
20
Filtered Thrust
Command
0
0 2 4 6 8 10 12
Time (s)
(c)
Figure 31: Thrust time histories for cold-flow closed-loop tests (a) ColdFlowCL12 (b) ColdFlowCL13 and
(c) ColdFlowCL14.
36 of 42
800
Thrust (N)
600
400
Raw Thrust
200 Filtered Thrust
Command
0
0 2 4 6 8 10 12
Time (s)
1000
800
Thrust (N)
600
400
Raw Thrust
200 Filtered Thrust
Command
0
0 2 4 6 8 10 12 14 16
Time (s)
Figure 33: Thrust signal compared to command for for Test HTPBCL2.
1000
800
Thrust (N)
600
400
Raw Thrust
200 Filtered Thrust
Command
0
0 2 4 6 8 10 12 14 16
Time (s)
The gains used for PressHTPBCL4 were meant to repeat the amount of damping as observed for HTP-
BCL1. Comparison of Figures 32 and 35 shows that this was the case. Gains were lowered for PressHTP-
BCL5 to reduce the amount of ringing in the controller response to the second step. Figures 35 and 36
show that the trade off for decreased ringing in the controller response is an increase in response time. The
balance between oscillations and response time can be adjusted with the gains to meet mission requirements.
37 of 42
1000
800
Thrust (N)
600
400
Raw Thrust
200 Filtered Thrust
Command
0
0 2 4 6 8 10 12 14
Time (s)
Downloaded by "Indian Institute of Technology, Bombay" on March 31, 2025 | http://arc.aiaa.org | DOI: 10.2514/6.2012-4200
1000
800
Thrust (N)
600
400
Raw Thrust
200 Filtered Thrust
Command
0
0 2 4 6 8 10 12 14
Time (s)
3. Simulator Comparison
Simulator results shown in Fig. 21 had the same operating conditions as PressHTPBCL4, results shown in
Fig. 35. As with the cold-flow controller, the actual response was slightly less damped than the simulator
response. This result was typical of both hot-flow controllers. Although the response was less damped, gain
sets predicted to be stable by the simulator were also stable in the controllers. There was no need to reduce
the gains as there was with the cold-flow controller.
38 of 42
0.68
HTPBCL2
0.66 PressHTPBCL3
Servo Input Voltage
0.64
0.62
Downloaded by "Indian Institute of Technology, Bombay" on March 31, 2025 | http://arc.aiaa.org | DOI: 10.2514/6.2012-4200
0.6
0.58
38
HTPBCL2
PressHTPBCL3
37
% MVT
36
35
34
6 6.2 6.4 6.6 6.8 7 7.2 7.4 7.6 7.8 8
Time (s)
(b)
Figure 37: Servo input voltage and position at beginning of ramp up command.
The thrust-feedback controller used in HTPBCL2 was able compensate for the eroded nozzle because
the thrust was controlled directly. Figure 31 shows that the thrust from HTPBCL2 was approximately the
same as the tests without nozzle erosion, while Fig. 38 shows that the control valve was opened substantially
wider to produce the same thrust. The chamber pressure-feedback test, PressHTPBCL5, on the other hand,
was unable to compensate for the change in the nominal relationship between thrust and chamber pressure
using the a priori conversion discussed in Subsection IV.B, even though the chamber pressure remained at
nominal levels. Because thrust cannot be measured directly in-flight, this problem could be encountered
with any in-flight feedback mechanism. However, shifts from nominal controller operating conditions can be
39 of 42
90
HTPBCL1
HTPBCL2
80
PressHTPBCL3
PressHTPBCL4
PressHTPBCL5
Downloaded by "Indian Institute of Technology, Bombay" on March 31, 2025 | http://arc.aiaa.org | DOI: 10.2514/6.2012-4200
70
dĞƐƚĚĂƚĂ
ĚĞŵŽŶƐƚƌĂƚŝŶŐ
60 ĞĨĨĞĐƚƐŽĨŶŽnjnjůĞ
% MVT
ĞƌŽƐŝŽŶ
50
40
30
20
0 2 4 6 8 10 12 14 16
Time (s)
Figure 38: Control valve position during hot-flow closed-loop throttling tests.
6. Steady-State Accuracy
Reducing the amount of motor-to-motor variability was the primary motivation for using closed-loop throt-
tling instead of open-loop. The first four seconds of each of the five hot-flow throttle tests presented here
had the same target thrust, 700 N. Steady-state controlled response was evaluated by calculating a mean
and standard deviation of the recorded thrust for all tests in the 2-4 second time interval, avoiding the initial
transients. The standard deviation of the thrust during the initial 700 N set point was 9.97 N for all five
controlled tests. If Test PressHTPBCL5, which used chamber-pressure feedback and had nozzle erosion, is
neglected this amount reduces to 3.34 N, nearly an order of magnitude lower than the standard deviation
of 30.12 N for the uncontrolled motor burns shown in Fig. 7. The mean thrust during the initial 700 N set
point for all five tests was 1.14% higher than the set point as measured on the cDAQ, or 0.53% higher if
PressHTPBCL5 is neglected.
VIII. Conclusion
Hybrid rocket throttling has potential applications both in space and for use in atmospheric flight. Open-
loop throttle testing established physical relationships between the control valve and motor response. Liquid
rocket engines are physically limited in deep throttle by an approximate 1.25 injector pressure ratio. This
limit was not observed in the hybrid rocket test motor configuration. Deep throttle testing showed that the
test motor configuration could produce stable combustion in a 66:1 turndown ratio and that the injector
pressure ratio increased with deeper throttling.
Closed-loop thrust-feedback throttle controllers were developed for a cold-flow nitrous oxide thruster and
for a nitrous oxide/HTPB hybrid rocket motor. A closed-loop chamber pressure-feedback throttle controller
was also developed for the nitrous oxide/HTPB hybrid rocket motor. Closed-loop cold-flow throttle testing
was performed before hot-flow testing began to evaluate the accuracy of the simulator. Slightly smaller
40 of 42
All throttle controller testing was performed using basic controllers and off-the-shelf equipment. Con-
troller performance can be improved using a more sophisticated throttle control system with features such as
a control valve actuator with reduced or no deadband, a control valve with a more linear flow profile, using
an automation controller capable of running higher loop frequencies, and an online, adaptable conversion
between thrust and the feedback mechanism. Even with all of the hardware limitations in this experiment,
the throttle control system substantially reduced the variability of the test motor configuration as compared
to the uncontrolled tests.
References
1 Chang, I.-S., “Investigation of Space Launch Vehicle Catastrophic Failure,” AIAA Paper 95-3128 , July 1995.
2 Safie, F. M. and Fox, E. P., “A Probabilistic Design Analysis Approach for Launch Systems,” 27th AIAA, SAE, ASME,
and ASEE, Joint Propulsion Conference, Vol. AIAA-1991-3372, 1991.
3 Maggie, G., “Space Shuttle Probabilistic Risk Assessment: Methodology and Application,” International Symposium on
Product Quality and Integrity, Proceedings of the Reliability and Maintainability Symposium, Las Vegas NV , 1996.
4 Anon., “Hazard Analysis of Commercial Space Transportation; Vol. 1: Operations, Vol. 2: Hazards, Vol. 3: Risk Analy-
and Human Services Public Health Service Agency for Toxic Substances and Disease Registry (ATSDR). Atlanta GA, 1997,
pp. 1–224.
6 DeSain, J. D., “Green Propulsion: Trends and Perspectives,” Crosslink, http://www.aero.org/publications/crosslink/summer2011/04.html,
Weapons Laboratory, Report AD-784 602, Kirtland AFB, New Mexico, July 1974.
9 Altman, D., “Hybrid Rocket Development History,” 27th AIAA, SAE, ASME, and ASEE, Joint Propulsion Conference,
1991.
10 Moore, G. E. and Berman, K., “A Solid-Liquid Rocket Propellant System,” Jet Propulsion, Vol. 26, No. 11, November
1975.
15 Ordahl, D., “Hybrid Propulsion,” Space Aeronautics, Vol. 41, No. 4, April 1964, pp. 108–113.
16 Hamers, J., “Experimental Investigation of Prepackaged Hybrid Propellant Systems,” Final Report AFRPL-TR-67-168 ,
1967.
17 French, J. R., “AMROC Industrial Launch Vehicle: A Low Cost Launch Vehicle,” SAE Technical Paper 871336 , 1987.
18 Kniffen, R., McKinney, B., and Estey, P., “Hybrid Rocket Development at the American Rocket Company,” AIAA-90-
2762 , 1990.
19 Estey, P. N. and Flittie, K. J., “Aquila - The Next Generation Launch Service for Small Satellites,” 14th International
Communication Satellite Systems Conference and Exhibit, Washington DC , March 22-24, 1992.
41 of 42
Motor,” AIAA/SAE/ASME/ASEE 29th Joint Propulsion Conference and Exhibit, Monterey, CA, June 28-30, 1993.
22 Flittie, K. J., Estey, P. N., and Kniffen, R., “The Aquila Launch Vehicle: A Hybrid Propulsion Space Booster,” Acta
hybrid motors under the joint government/industry IR&D program,” AIAA Paper 93-2552 , 1993.
25 Carpenter, R. L., Boardman, T. A., Claflin, S. E., and Harwell, R. J., “Hybrid propulsion for launch vehicle boosters: A
program status update,” 31st AIAA/ASME/SAE/ASEE Joint Propulsion Conference and Exhibit, San Diego CA, July 10-12,
1995.
26 Arves, J., Gnau, M., Joiner, K., Kearney, D., McNeal, C., and Murbach, M., “Overview of the hybrid sounding rocket
“Design and Development of a 100 km Nitrous Oxide/Paraffin Hybrid Rocket Vehicle,” AIAA 2007-5362 , 2007.
29 Doran, E., Dyer, J., Marzona, M. T., Karabeyoglu, A., Zilliac, G., Mosher, R., and Cantwell, B., “Status Update Report
for the Peregrine Sounding Rocket Project: Part III,” AIAA 2009-4840 , 2009.
30 Austin, B., Heister, S., Dambach, E., Wernimont, E., and Meyer, S., “Variable Thrust, Multiple Start Hybrid Motor
Solutions for Missile and Space Applications,” 46th AIAA/SME/SAE/ASEE Joint Propulsion Conference, Nashville TN , July
25-28, 2010.
31 Eilers, S. D., Wilson, M., and Whitmore, S., “Analytical and Experimental Evaluation of Aerodynamic Thrust Vectoring
on an Aerospike Nozzle,” 46th AIAA/ASME/SAE/ASEE Joint Propulsion Conference and Exhibit, Nashville, TN , 2010.
32 Eilers, S. D., WIlson, M. W., Whitmore, S., and Peterson, Z., “Side Force Amplification on an Aerodynamicclly Thrust
Hybrid Rocket Fuels,” 47th AIAA/ASME/SAE/ASEE Joint Propulsion Conference & Exhibit, 2011.
34 Anon., “Liquid Rocket Valve Assemblies,” NASA-SP-8097 , 1973.
35 Anon., Control Valve Handbook , Fisher Controls International LLC, 2005.
36 Anon., “Flow Equations for Sizing Control Valves,” ANSI/ISA-75.01.01 (IEC 60534-2-1 Mod)-2007 , Nov. 2007.
37 “V-Control Ball Valve Selection Guide Technical Bulletin No. 1006,” Tech. rep., Flow-Tek, Sept. 2008.
38 Miller, R. W., Flow Measurement Engineering Handbook, 2nd Ed., McGraw-Hill Publishing Company, 1989.
39 Anon., “Hydroxyl Terminated Polybutadiene Resins and Derivatives,” Sartomer Product Bulletin, March 2006.
40 Anon., “PAPI 94 Product Information,” Dow Plastics Form No. 109-00707-801XQRP , August 2001.
41 Karabeyoglu, A., Stevens, J., and Cantwell, B., “Investigation of Feed System Coupled Low Frequency Combustion
for Nonpolar and Polar Fluids,” International Journal of Thermophysics, Vol. 24, No. 1, 2003, pp. 1–39.
44 Span, R. and Wagner, W., “Equations of State for Technical Applications. II. Results for Nonpolar Fluids,” International
and Applications II. Users Manual and Program Description,” Tech. rep., NASA RP-1311, 1996.
48 Carmicino, C., “Acoustics, Vortex Shedding, and Low-Frequency Dynamics interaction in an Unstable Hybrid Rocket,”
Joint Propulsion Conference & Exhibit, Seattle WA, July 6-9, 1997.
54 Giuliano, V. J., Leonard, T. G., and Lyda, R. T., “CECE: Expanding the Envelope of Deep Throttling Technology
in Liquid Oxygen/Liquid Hydrogen Rocket Engines for NASA Exploration Missions,” 46th AIAA/ASME/SAE/ASEE Joint
Propulsion Conference & Exhibit, July 2010.
55 Betts, E. M. and Frederick, Robert A., J., “A Historical Systems Study of Liquid Rocket Engine Throttling Capabilities,”
46th AIAA/ASME/SAE/ASEE Joint Propulsion Conference & Exhibit, Nashville TN , July 25-28, 2010.
42 of 42