Untitled
Untitled
Thesis
Submitted to
UNIVERSITY OF DAYTON
The Degree
By
UNIVERSITY OF DAYTON
Dayton, Ohio
December, 2014
EXPERIMENTS WITH A HIGH PRESSURE WELL STIRRED REACTOR
APPROVED BY:
________________________________ ________________________________
Dr. Scott Stouffer, Ph.D. Dr. Jamie Ervin, Ph.D.
Advisory Committee Chairman Committee Member
Senior Research Engineer
Energy and Environmental Engineering
________________________________
Dr. Alejandro Briones, Ph.D.
Committee Member
Senior Research Engineer
University of Dayton Research Institute
________________________________ ________________________________
John G. Weber, Ph.D. Eddy M. Rojas, Ph.D., M.A., P.E.
Associate Dean Dean
School of Engineering School of Engineering
ii
ABSTRACT
The existing WSR has been successfully used to evaluate pollutant emissions,
pressure. The HPWSR project extends the capabilities of this laboratory tool, allowing
the generation of benchmark quality data over a range of operating pressures. Lean blow
out data collected is correlated against loading parameter, forming fuel stability loops.
Combustion stability of JP-8 will be evaluated over a range of operating pressures from 4
higher pressures than with previous hardware. Lean emissions will be collected
simultaneously and used to further characterize the HPWSR operability and lean
combustion performance. This work will validate nitrogen dilution using true low
pressure testing compared to dilution in the same configuration, and further extend WSR
iii
ACKNOWLEDGEMENTS
Scott Stouffer, thank you for your support and guidance in making this research
possible. Our time spent guiding me through the engineering design process has helped
me to develop the skills and connections that will lead me to a truly rewarding career
To Craig Neuroth, Brad Day and David Baker, thank you for your support and
effort in integrating our hardware to the high pressure facility. Without your help and
To Alejandro Briones and Jamie Ervin, thank you for participating in my thesis
defense and for your time reviewing my thesis. Your comments helped strengthen the
To my wife, Kerianne, thank you for your support as I went through for school,
and for continuing to push me to finish this thesis even when I would have rather been
watching football.
Finally, a special thanks to funding from the Hons von Ohain Fellowship, which
iv
TABLE OF CONTENTS
ACKNOWLEDGEMENTS ............................................................................................... iv
v
2.4.1. Emissions and Soot Formation................................................................ 16
vi
3.6.2. High Pressure Operation ......................................................................... 61
TECHNIQUE .............................................................................................. 62
CHAPTER 5 CONCLUSION.......................................................................................... 87
5.2. CONCLUSIONS......................................................................................... 88
BIBLIOGRAPHY ............................................................................................................. 92
vii
A.3. PROBE MEASUREMENTS WITH RADIATION-CONDUCTION
viii
LIST OF FIGURES
Figure 2.3: Lower half of the toroid with the jet ring in the WSR housing ...................... 15
Figure 2.4: Set-up and instrumentation points of the ceramic total temperature probe .... 22
Figure 3.8: Mixture temperature vs. dewpoint at maximum vaporizer pressure .............. 43
Figure 3.12: Flow path schematic for experiments with an external vaporizer ................ 59
Figure 3.13: Laboratory set-up of the ceramic total temperature probe ........................... 65
ix
Figure 4.1: Lean blowout of alternative fuels in a metal reactor ...................................... 71
Figure 4.7: Blowout points taken with internal and external vaporization ....................... 78
Figure 4.13: JP-8 combustion stability curve using simulated low pressure .................... 84
Figure 4.14: Correlation of true low pressure to nitrogen dilution data ........................... 86
Figure A.4: Comparison of temperatures in an ethylene Hencken burner flame ........... 100
Figure A.5: Intrusive impact of a probe on a Hencken burner flame ............................. 101
Figure A.6: Comparison of corrected probe hydrogen flame temperatures ................... 103
Figure A.7: Comparison of corrected probe ethylene flame temperatures ..................... 104
x
Figure A.8:WSR exhaust temperature with probe radiation-conduction correction ...... 105
xi
LIST OF TABLES
Table 3.5: Test Conditions for Sub-Atmospheric Blowout with External Vaporization .. 61
Table 3.6: Test Conditions for High Pressure Blowout with External Vaporization ....... 62
Table 3.7: Test Conditions for High Pressure Blowout with Internal Vaporization ........ 63
Table 3.8: Test Conditions for Sub-Atmospheric Blowout with Internal Vaporization ... 64
Table 3.9: Test Conditions for Ceramic Probe to Thermocouple Comparison ................ 67
xii
LIST OF SYMBOLS AND ABBREVIATIONS
English Symbols
A = Arrhenius reaction rate expression
C = constant of integration
D = diameter
E = activation energy
h = heat transfer coefficient (W/m2-K)
k = thermal conductivity (W/m2-K)
= mass flow rate (lb/min)
Greek Symbols
γ = ratio of specific heats
ρ = density (kg/m2)
τ = residence time time (s)
Φ = equivalence ratio
xiii
Subscripts
1 = location in flame
2 = downstream location
4 = turbine inlet
∞ = freestream
cond = conduction
condi = condition
conv = convection
rad = radiation
stoich = stoichiometric
t = total
x = location
Abbreviations
AFRL = Air Force Research Laboratory
HPCRF = High Pressure Combustion Research Facility
SLPM = Standard Liters Per Minute
WSR = Well Stirred Reactor
xiv
CHAPTER 1
INTRODUCTION
both fuel consumption and pollutant emissions. Growing concerns over the supply of oil
currently used to make transportation fuel has spurred the development of alternative
fuels that could be suitable for gas turbine engines. These alternative fuels seek to use
commonly available feed stock to produce viable transportation fuels to help reduce the
reduce fuel consumption. As the fuel to the combustor is reduced, the flame is more
highly stressed. In order for next generation combustors to optimize efficiency and
operate in the most fuel-lean conditions, a complete understanding of the stability limits
physical and chemical properties is required to determine if the new fuel is feasible in a
gas turbine combustor, and if it can meet the emissions and stability requirements of
modern engines. All current gas turbine engines are designed to operate on oil-based jet
fuel. Testing of alternative fuels is needed to demonstrate that their combustion is stable,
generates adequate power and burns cleanly to make alternative fuels viable replacements
1
Research of fundamental combustion characteristics can be achieved in a Well
Stirred Reactor (WSR). The well stirred reactor is a laboratory idealization of a gas
turbine combustor that utilizes pre-mixed, pre-vaporized fuel and air flow to isolate
chemical kinetic effects from the fluid dynamic effects. The WSR is primarily used to
evaluate particulate and pollutant emissions and the combustion stability of hydrocarbon
fuels.
1.2. OBJECTIVES
In a real gas turbine engine, the combustor is operating at elevated pressures and
states that increasing the temperature and pressure of the combustion gases improvers the
atomization of fuel droplets, increases the turbulent flame speed, and increases the
reaction rate of combustion. The results is that increased temperature and pressure yield
more efficient combustion that produces less pollutants and is more stable than
could also exhibit sub-atmospheric pressures if the combustor were to blow out while at
cruising altitude [21, 23]. It is critical to the safety of the plane that the stability of
The objective of the current study is to use the existing geometry of the WSR and
adapt it for experiments at elevated and sub-atmospheric pressures that are more
representative of the conditions present in a gas turbine engine. The expanded capability
of the WSR will allow characterization of the effects of increased pressure on the
2
combustion stability and efficiency of the flame at higher pressures. Additionally, the
capability of the WSR to achieve sub-atmospheric pressure will allow the study of the
Chapter 2 elaborates on the background of the WSR, and the history of the design and
past experiments using the geometry. The section also contains explanation of important
nitrogen dilution.
Chapter 3 details the design and experimental setup and procedure of the high pressure
WSR. Discussion includes many of the design challenges when dealing with fuel
Chapter 4 presents and discusses the experimental results of studies using the WSR at
elevated and subatmospheric pressure as well as the lean blowout stability of alternative
Chapter 5 presents the conclusions, and suggests possibilities for future study.
3
CHAPTER 2
BACKGROUND
operating the combustor at fuel-lean conditions. As the fuel flow in decreased, however,
the fuel approaches its lean flammability limit where combustion can no longer be
sustained. The equivalence ratio at which combustion ceases is referred as lean blowout.
Operating an engine near lean blowout, however, requires a solid understanding of the
flame stability over the range of air flow rates, pressures and temperatures that will be
experienced during flight. Therefore, it is important that the stability characteristics of the
combustor be defined so that all operating conditions can be shown to fall within the
Often in engine testing, the blowout equivalence ratio is plotted against the chamber
reference velocity, due to the relationship between bulk flow velocity and flame speed.
[21]
If the velocity of the gas stream through the combustor is higher than the turbulent
flame speed, the flame will move downstream and the combustion will not be stable. [21]
However, plotting stability loops in this way requires that a new curve be created for each
4
different pressure. Figure 2.1 shows typical stability curves as a function of combustor
The stability curve will continue to increase in fuel/air ratio as combustor velocity
increases until the flame reaches the stoichiometric fuel/air ratio. The stoichiometric
fuel/air ratio is the most stable condition for combustion because it is the condition of
highest heat release, and it represents the peak on the stability curve. Any more fuel
added to the system past the stoichiometric fuel/air ratio will decrease the combustion
temperature, and therefore the stability. In addition to the fuel/air mixture, the combustor
operating pressure impacts the combustion stability. As the pressure is increased, the
flame speed and reaction rate will increase, improving the stability of the flame by
allowing the flame front to exist in a higher velocity air stream. For each different
Early work performed by Longwell and Weiss [41] demonstrated that flame
stability could be correlated with the air loading parameter. The air loading parameter is
5
(1)
In Equation 1, n is the effective order of the reaction, is the combustor mass flow rate,
V is the reactor volume and P is the reaction pressure. Longwell and Weiss [41] determined
recommended that the reaction order can be related to the combustor equivalence ratio,
. The correlations by Kretschmer and Odgers[4] showed that for lean mixtures
Figure 2.2 shows similar stability curves correlated with the air loading
parameter. The benefit of using the loading parameter is that the x-axis is a function of
pressure and mass flow, allowing data collected at different pressures, combustor
6
In Figure 2.2, many different operating points are correlated to a single curve,
increase the loading parameter, the mass flow into the combustor can be increased.
However, only so much mass flow may be passed for a given volume. As the mass flow
through the reactor is increased, the residence time of the fluid in the reactor volume is
decreased [21]. Because the volume of the burner is fixed, decreased residence time
increases the fluid velocity. As shown in Figure 2.1, the stability of a flame decreases
with increased velocity. To continue increasing the mass flow rate, the combustor volume
can be increased to maintain the flame in a stable burning range. Additionally, from the
definition of loading parameter, the operating pressure can be lowered to increase the
loading parameter. In general, lowering the operating pressure decreases the range of
operating conditions over which the combustor can be operated. Lower operating
pressures decrease the turbulent flame speed [39], therefore high mass flow rates lead to
flame blowout. All of these combinations are represented in a stability loop plotted
against loading parameter, making the loading parameter a valuable tool for combustor
characterization.
Many studies have been conducted using various laboratory combustors that
explore the stability characteristics of fuel at atmospheric pressures. Ballal et al. [5]
conducted a combustion stability experiment using a research reactor that simulated the
main features of an annular gas turbine combustor, with propane as a fuel. Lean blowout
was measured as a function of air loading parameter. Experimental data were compared
to predictions from the Perfectly Stirred Reactor (PSR) theory. Ballal et al.[5]
7
specific series of events which indicate that the LBO mechanism is very complex
behaves like a PSR over an air loading parameter range of 0.1 – 100
Furthermore, the results illustrated that blowout data may be correlated with the air
loading parameter, as derived from PSR theory due to the fact that the loading parameter
accounts for the effects of residence time and flow velocity as well as pressure.
Extensive testing is performed to ensure that the flame is statically stable over a
large range of operating conditions, ensuring a safe flight. When engaging in combustion
stability testing, it is important to evaluate the system at the extremes of the operating
range. One limitation observed in the previous studies is that combustion at atmospheric
pressure does not yield a significant variation in blowout equivalence ratio (when
correlated with loading parameter) to construct a full stability loop. When combustion
stability is plotted against the air loading parameter, as shown in Figure 2.2, the variation
in the range of conditions is dependent upon the mass flow rate and operating pressure.
Atmospheric testing in a WSR operates at a fixed pressure and velocity, thus limiting
mass flow rate and therefore the loading parameter. In a combustion system, the worst
fuel stability is fully exposed at the lowest combustion pressures [1]. In aviation, low
operating pressure occurs at high altitude due to the decrease in atmospheric pressure as
altitude is increased. [23] To understand low pressure combustion, the most highly
environment. To mitigate the high cost and complexity of full scale, true low pressure
8
combustion test, and several methods have been proposed to simulate low reaction
pressure.
A study by Lefebvre and Halls [1] used a water injection technique to simulate low
combustion pressure and increase the loading parameter of the combustor. In a practical
gas turbine, as altitude increases, the pressure decreases, and, correspondingly, the heat
release rate of the flame. The addition of water vapor into the system lowers the
combustion temperature and thereby simulates the lower heat release of low pressure
flames. In their study, fuel and air flow rates were held constant, and water was gradually
mixed with the fuel until extinction occurred. The process was repeated over a range of
flow rates, allowing a full stability loop to be constructed. Lefevbre and Halls concluded
that the water injection technique allows the point of peak stability to be clearly
defined.[1]
Additional work in the simulation of low pressure, high loading parameter flames
was conducted by Sturgess et al. [10] The study proposed that a simulation of low pressure
via nitrogen dilution was achieved when the reaction rates were matched. The steps for
the simulation proposed by Sturgess et al. [10] is shown below. The simulation began with
a bimolecular global reaction for a propane-air flame with excess nitrogen, given by,
(2)
(3)
9
K is the ratio of the mass flow of the excess nitrogen to the fuel flow rate at a
given condition. Atmospheric nitrogen will always be present in the combustion gas
when using air, however, when K is greater than 0, additional nitrogen has been added to
the flow as a diluent. When K = 0, the reactants are solely propane and air. An Arrhenius
reaction rate expression for a bimolecular single step reaction, proposed by Longwell et
(4)
Equation 4 can be rearranged in terms of air loading parameter and takes the
form,
(5)
Equation 5 can be used to represent the reaction rate for a propane-air flame with
rate simulation is achieved when the simulated loading parameter is equal to the desired
(6)
In Equation 6, R is the reaction rate expression from the right hand side of
Equation 4 for the desired case. In the simulation, the pressure is 1 atm and excess
nitrogen is present (i.e., K > 0) and in the desired case, the pressure is subatmospheric
Sturgess et al. [10] used a propane fueled diffusion flame combustor operating at
constant air and fuel flow rate to test the effect of excess nitrogen on lean blowout
10
characteristics. Nitrogen was added into the air incrementally until lean blowout was
reached. Sturgess et al. [10] showed that the nitrogen dilution technique allowed the air
loading parameter to be increased over three orders of magnitude. Additionally, the data
Currently, the rising cost of conventional oil derived fuels as well as concerns
over the depleting oil reserve have led to rapidly expanding research into alternative
energy sources. Of main concern to the aerospace industry is the development of gas
turbine fuels from renewable sources. Commercial aviation is the largest consumer of
aviation fuel, with the US accounting for 44% of the global demand.[7] Recent efforts
have targeted development of a jet fuel derived from a renewable source. Potential JP-8
substitutes can be derived from feed stock of coal, natural gas or various biomass sources
such as algae, wood pulp or commercial fat, oil and grease waste.[7] A paper by Blakey et
al. [35] in 2011 detailed the process used to create the currently available biofuels.
hydro-treating of a bio mass feed stock. In the Fischer-Tropsch process, coal is gasified,
and turned into syngas, a mixture of hydrogen and carbon monoxide. The syngas passes
hydrocarbon fuel with properties similar to JP-8. Hydro-treated renewable jet fuel (HRJ)
unsaturated hydrocarbon chains. Blakey et al. [35] showed that in a life cycle analysis,
biomass oils such as camelina, palm and soy produce less CO2 emissions per MJ of
11
2.2. PREVIOUS WSR CONFIGURATIONS
The WSR has been used for many years to study combustion reactions. A highly
turbulent premixed air fuel mixture is injected into the reactor volume, continuously
mixing the combustion products with incoming reactants. The high degree of mixing
results in near uniform temperature and species profiles across the reactor volume,
reactor walls are constructed from ceramic to allow the reactor to operate in a near
adiabatic state. The combination of the high rate of mixing and adiabatic nature of
ceramic test sections yields a reactor that is ideal for testing the purely kinetic aspect of
fuels, de-coupling fluid dynamic phenomena. The WSR removes the fluid mechanic
allowing evaluation solely of the fuel oxidation chemistry in question. Sections 2.2.1-
2.2.3 discusses some of the WSR designs that were used in the past, as well as an
2.2.1. Spherical
In 1955, Longwell and Weiss [41] designed the first WSR to study high
temperature kinetics of hydrocarbon fuels. The reactor used by Longwell and Weiss was
made of two spherical halves of insulating firebrick. The upper temperature limit of the
firebrick was 1900 K. The reactor was enclosed in a nickel test section that allowed
instrumentation that included optical access, emissions probes and thermocouples. The
combustion products exhausted through holes drilled in the reactor. The work from
12
Longwell and Weiss using the spherical WSR suggested that lean blowout data may be
correlated to the loading parameter, derived from the kinetic theory. Longwell and Weiss
assumed an overall reaction order (n) to be equal to 1.8 for iso-octane. The data
One of the main disadvantages of the spherical reactor design was the material
properties of the firebrick. The high thermal conductivity resulted in high heat loss
through the walls, and the material quickly degraded when exposed to the high
In 1984, Nenniger et al. designed the toroidal WSR. [27] The reactor had the same
internal volume (250 mL) as the spherical reactor used by Longwell and Weiss. [41] The
toroidal reactor was designed to achieve a high rate of mixing. 42 inlet jets were angled
20° off the radius to promote jet penetration and encourage bulk swirl tangential to the
toroid. Nenniger et al. used a tracer dye and laser-induced fluorescence (LIF) to visualize
was used and improved by Zelina [42] then Stouffer et al. [33,36] The reactor is composed of
two half-toroid sections forming a reactor volume of 250 mL. Several ceramic
compounds have been used in previous tests, including alumina, fused silica, and zirconia
ceramic. The reactor halves sandwich an Inconel jet ring. The jet ring consists of 48 jets
drilled at 20° off the radius. The offset promotes swirl in the reactor and increases the
13
turbulence of the fluid. High turbulence increases molecular level mixing of the fuel and
air, establishing the well stirred condition. The jet ring is fed to tubes through which the
fuel/air mixture is introduced to the reactor. Additionally, the jets are sized such that the
passages are aerodynamically choked at the flow rates encountered during testing.
Choking the jets accomplishes two things. First, the high velocity of the jets enhances the
mixing in the reactor, and second chocking isolates the air supply from downstream
( , the reactor pressure will be less than 52.8% of the upstream pressure if the jet
ring is choked.
The fuel-air mixture is reacted in the toroid and exits through a single exhaust port
in the center of the reactor volume. Due to the swirl induced by the jets, a flow
straightener is placed in the exhaust port to avoid entrainment of outside air into the
The lower half of the reactor is instrumented through four holes. The typical
pressure tap to measure reactor pressure, an emission probe and an igniter. In several
studies, a stepper motor was used to sweep the thermocouple through the reactor,
generating a temperature profile along the minor radius of the toroid. The igniter uses
spark ignition and is fed from an aircraft exciter box. The igniter is recessed from the
main flow to preserve the material. Other designs use a pneumatic actuator to retract the
igniter form the flow after ignition is established. The pressure tap is made from ceramic
and sits flush with the top of the reactor. The location of the typical instrumentation is
14
called out. The WSR setup illustrated in Figure 2.3 was used in the studies by Zelina, [42]
Figure 2.3: Lower half of the toroid with the jet ring in the WSR housing [2,28]
temperature. High jet ring temperatures can be caused by inadequate cooling by the
gases to escape the test section. In several studies, supplemental jet ring cooling was
achieved through a cooling ring that allowed nitrogen flow to cool the exterior jet ring.
The nitrogen provided cooling and the inert environment surrounding the reactor,
The WSR has been used extensively over the years to evaluate fundamental
combustion performance. Specifically, the WSR has been used to evaluate the gaseous
15
emissions, soot formation and combustion stability. The following section details some of
grow, the importance of understanding and optimizing the combustion process grows as
well. The well stirred reactor is used extensively in the evaluation of emissions
A 1994 study by Zelina et al. [13] used a WSR to evaluate the emissions of CO,
CO2, NOx and unburned hydrocarbons (UHC) for ethane, methane and a hydrocarbon
blend. Zelina et al.[13] showed that the CO emissions varied little among the three fuels
dramatically as the equivalence ratio was increased. Similarly, the UHC emissions
showed little variation between the three fuels, reaching minima at . The UHC
emissions increased in both leaner and richer mixtures due to decreasing reaction
temperatures. NOx is strongly temperature dependent, and Zelina et al. [13] showed a peak
NOx concentration at an equivalence ratio near unity, where the combustion temperatures
are highest. NOx emissions were highest (in descending order) for the HC blend, ethane
and methane. The NOx concentrations correspond with the flame temperature as well,
with the HC blend being the hottest, and methane the coldest. The higher NOx was
formed when buring the HC blend due to the higher reaction temperatures increasing the
thermal NOx production. Zelina and Ballal [14] also calculated the combustion efficiency
for these fuels, using the WSR. Emissions of CO, NOx and UHC display the inefficiency
of the combustion process, and are, therefore, useful for calculating the combustion
16
efficiency. Zelina and Ballal [14] showed that a ceramic WSR had a combustion efficiency
of greater than 99% for all fuel-lean equivalence ratios up to stoichiometric. For rich
the production of CO and UHC are increased due to lower equilibrium flame
temperatures, lack of oxygen and surplus of fuel. Zelina and Ballal showed little variation
In 1997, Blust and Ballal [11] investigated the emissions characteristics of liquid
and gaseous fuels using a WSR. While gaseous fuels can provide a basis for modeling
and fundamental combustion research, liquid fuels are used almost exclusively in
blend were used in the experiment, and the emission index of CO and UHC was
measured. The emission index (EI) is the mass of the species of interest divided by the
mass of fuel taken to produce the sample. Blust and Ballal [11] showed that alkanes (with
Additionally, it was shown that increasing molecular weight between alkanes (such as n-
heptane and n-dodecane) did not have an effect on CO production. The aromatics
produced less CO as the CN was increased. Finally, it was shown that methane produced
the most CO of all fuels tested, mainly due to its low molecular weight in the EI
calculation. In addition, methane produced the most UHC, followed by toluene. The
UHC production of alkanes rapidly decreased with increasing CN. Aromatics produced
less UHC than alkanes, with the exception of ethylbenzene, which behaved similar to the
17
Additional studies have been performed using a WSR to characterize soot
formation. While CO and UHC directly correlate to combustion efficiency and are
study in 2005 by Stouffer et al. [33] used a WSR to investigate methods for soot reduction.
cyclohexane, pyridine and quinoline. Collected carbon mass was compared to fuel
without additives. Stouffer et al. explored the effect of chemistry and temperature on fuel
additives. The inlet temperature of the baseline fuel was adjusted to account for the
nitroalkanes showed a reduction in carbon mass across all equivalence ratios tested, with
4% showing the largest reduction. Nitroalkanes were the most effective at soot reduction
of all the additives tested; however, they were accompanied by an increase in NOx
emission. Cyclohexane decreased soot production over the entire range of conditions.
Stouffer et al. concluded that the soot reduction on the lower end of the equivalence ratio
range was chemical, while reduction at higher equivalence ratios was due to increased
flame temperature. The least effective additive used by Stouffer et al. in the study was
quinolone, which increased the carbon mass over all of the test conditions. [33]
Many studies have been conducted using various laboratory combustors that
explore the stability characteristics of fuel at atmospheric pressures. Ballal et al. [5]
conducted a combustion stability experiment using a research reactor that simulated the
main features of an annular gas turbine combustor, and used propane as a fuel. Lean
18
blowout was measured as a function of air loading parameter. Experimental data were
compared to prediction from Perfectly Stirred Reactor (PSR) theory. Ballal et al. [5]
systematic and detailed series of events which indicate that the LBO mechanism is very
complex. Comparison of experimental data showed that the research combustor behaved
like a PSR over an air loading parameter range of 0.1 – 100 . [5]
A study by Stouffer et al. [36] used a well stirred reactor (WSR) to evaluate the
lean blowout characteristics of standard JP8 and a synthetic fuel derived from the Fisher-
Tropsch process. Data collected for both JP8 and Fisher-Tropsch fuels were correlated
against air loading parameter and showed a 2.8% difference in the blowout equivalence
ratio between the two fuels. A full combustion stability loop could not be created, and the
blowout equivalence ratio varied on 1% as the loading parameter was doubled. The
In 2012, Karalus et al. [25] conducted a lean blowout study with premixed
hydrogen-air flame. The study aimed to characterize the mechanism of lean blowout in a
premixed flame, and compared the experimental results to CFD predictions. Karalus et al.
showed that lean blowout did not occur when the flame was spatially homogeneous. The
flame was stabilized by a recirculation zone of combustion products near the base of the
inlet jet. Additionally, Karalus et al. [25] achieved lean blowout by maintaining air flow
rate constant while lowering fuel flow. The study showed that this method to achieve lean
blowout maintained near constant turbulence intensity parameter as lean blowout was
approached. Each reduction in the fuel flow was followed by a 15 minute wait time to
allow the reactor to achieve thermal equilibrium. Karalus et al.[25] stated that if lean
19
blowout is approached too rapidly, the hotter reactor walls will temporarily support
kinetics. [39] The majority of the flame temperature measurement is done with
calibration and installation, while providing reasonable accuracy and response time. The
main limitation, however, is the maximum temperature that a thermocouple can withhold.
For near stoichiometric combustion, the flame temperature can exceed 4000 °F. For lean
blowout studies where the equivalence ratio is maintained at or below approximately 0.6,
the flame temperature can reach 2600-2800 °F. [39] A type K thermocouple uses nickel-
chromium and nickel-aluminum leads and has a maximum temperature of 2285 °F. [2]
The low maximum service temperature makes a type K thermocouple unsuitable for most
metal temperatures and hardware health. Type B thermocouples are routinely used for
thermocouple is 3092 °F. [42] The high temperature capability of the type B thermocouple
20
conditions. However, a type B thermocouple cannot be used for all conditions,
possible. In high velocity conditions, the thermocouple can easily break or otherwise no
longer provide accurate readings. The issue of fragility and longevity is especially
To increase the strength of the probe for withstanding high flow rates in
technique as well as some non-intrusive techniques used in the combustion field. These
same non-intrusive techniques are used to evaluate the performance of the new
temperature probe.
The current study aims to develop and characterize a diagnostic tool that can be used
previously accessible only through laser diagnostics. Additionally, very little literature
exists detailing the quantitative effects of introducing a probe into a flow stream. In
21
combustion, thermocouples and emissions probes are used extensively. The current study
acquires laser measurements simultaneously with the probe to show the perturbations that
intrusive diagnostics impart to a flame, shedding light into the discrepancy between
measure pressure, temperature and emissions in combustors that do not have optical
access. Moreover, the hardware required to support the probe is common and the probe
will allow deployment in the field without the cost and complexity of laser diagnostics.
The ceramic temperature probe uses flow continuity and the isentropic gas
relations to calculate a flame temperature from gathered temperature and pressure data
across a series of orifices. Figure 2.4 illustrates the location and numbering of the stations
Sample to
Vacuum
Pump
Figure 0.5: Set-up and instrumentation points of the Total Temperature Probe
Figure 2.4: Set-up and instrumentation points of the ceramic total temperature
probe
22
Mass continuity states that the flow rate entering the system must be the same as that
(7)
Since the gas flowing through the system is combustion products well above the
Equation 7 can be rewritten using the ideal gas law, [23] shown below in Equation 8.
(8)
In Equation 7, it is necessary to know the area at the fluid entry and exit planes, the fluid
velocity, pressure and temperature. The pressure can be measured at a pressure tap in the
The velocity of the fluid may be difficult to determine for flow through a tube, however,
if the system consists of two orifices, a sufficient pressure drop will provide aerodynamic
choking. Therefore, the velocity is known to be the local speed of sound, [23] and
(9)
(10)
In Equation 9, station 1 and 2 represent the orifice in the flame and the downstream
orifice, respectively. A thermocouple inserted into the flow at the downstream orifice can
be used to measure the temperature at station 2, while the temperature at station 1 is the
flame temperature, and is the value of interest. The gas constant, R, is assumed to be the
same between the two orifices in the system, and therefore cancels out from the equation.
23
The composition of the products is assumed to be frozen once it enters the probe.
However, Cp, and, therefore, will vary with temperature. Since the area of the orifices
used is known, the final parameter to define is the fluid temperature. With the above
assumptions, Equation 9 can now be solved for the temperature at the orifice in the flame.
(11)
The flame total temperature is determined using the isentropic relations. Both the
(12)
(13)
Finally, it is assumed that the static and total temperatures are equivalent at each
station in the probe, due to the low flow velocities. A large probe with small orifice will
result in sonic velocities through the orifice, but the fluid will quickly expand in the body
of the probe and slow such that the static and total temperatures can be assumed equal.
The total temperature is within 99% of the static temperature for Mach numbers less than
0.15. The Mach number in the probe can be readily calculated from isentropic relations,
knowing the ratio between the probe internal area and the orifice area, . [23]
When applied to a real world system, the geometric area is not representative of
the flow area. Often, the area is qualified with a discharge coefficient. During the
development of the temperature probe, the discharge coefficient of the orifice in the
probe could not be directly measured. Instead, the probe orifice and downstream orifice
24
were flowed together to generate an effective area. With the geometric areas known, the
ratio of the discharge coefficients could be backed out. This ratio of the discharge
coefficients in the system was included in the calculation of the flow temperature.
error in the reading from the impact of the probe on the flow. The metal thermocouple
theory would also suggest that since the base of the thermocouple is cooler than the
flame, heat is removed from the measurement point through conduction. If an accurate
measurement is desired, it is necessary to correct for the heat transfer losses. Fortunately,
the main losses are through radiation and conduction. A study by West and Westwater[40]
treats a thermocouple as a fin and applies a correction for heat transfer losses. They
compared the results of a combined heat transfer correlation with temperature probe
radiation and conduction. Their results showed a reasonable correlation. Thus, the current
and connected to the sampling lines with a series of fitting and adapters. The cantilever
nature of the probe coupled with the constant cross-sectional area allows the probe to be
modeled as a fin. Additionally, the length of the probe, the low fluid flow rate through the
system and the number of stainless steel fittings used allow the assumption that the base
25
of the probe is at the wall temperature and will be relatively independent of flame
The heat equation can be applied to a differential element along the length of the
probe.
(14)
The heat transfer terms shown in Equation 14 can be expanded with detailed heat
flows, shown in Equation 15, where FA and Fe are the radiation angle factor and radiation
(15)
(16)
the following:
(17)
(18)
26
Equation 18 can be directly integrated if the probe has a constant perimeter and cross-
(19)
Where
Boundary conditions must be applied to solve for the constants in Equation 19. As
previously mentioned, it is assumed that the base of the probe is equal to the wall
The probe is used to measure the temperature at the choked orifice. Since the orifice is
significantly smaller than the diameter of the probe, the heat transfer across the orifice
negligible. Therefore, the assumption of no heat flux at the orifice is valid. Applying the
(20)
(21)
where . The value of the constants can be substituted into Equation 19,
resulting in a general equation for the temperature at any point, x along the probe.
However, the point of interest is the orifice in the flame, which occurs at . Applying
(22)
27
In Equation 22, is the actual gas temperature, is the wall temperature and is the
Laser diagnostic measurements allow a less intrusive method for the evaluation of
(i.e., emission probe, thermocouple, etc.,) the flame is disturbed. The perturbation caused
in the flame by the measurement technique has the potential to skew the accuracy of the
measurement, especially if a large probe is used to sample from a small flame. Fluid
temperature and species concentrations are the main parameters that are measured with
tune the laser over the desired characteristic absorption lines in the flame. When the laser
impacts a species of the same emission wavelength, a reduction in the signal intensity is
recorded. The intensity of the beam through the flame can be correlated to the species
concentration through the Beer-Lambert law [17]. The flame temperature is measured by
sweeping the absorption spectra over two different absorption lines for the same species.
The integrated absorbance of the selected species is a function solely of the flame
temperature. [2]
(CARS), which uses three laser beams, viz., a pump, a Stokes and a probe beam. The
optical signal generated by the interaction between the three beams and the sample is at
the anti-Stokes frequency, or the sum of the frequencies of the original three beams. If the
28
difference between the pump beam and the Stokes beam is equivalent to the Raman
The strength of the optical signal generated in with CARS diagnostics is a function of the
species. The number of excited species in a given state is a function of the temperature-
dependent Boltzmann distribution. Therefore, the strength of the signal may be correlated
by Hancock et al. [32] measured the temperature of various flames over a Hencken burner
with CARS. The Hencken burner is a small, laboratory flat flame burner that produces
consistent profiles suitable for research using a sensitive method such as CARS.
Hydrogen-air and ethylene-air flames were studied. Hancock et al. showed that the
temperature of the Hencken burner flame matched closely with the equilibrium flame
temperature, and that the flame temperature was uniform across the width of the flame
front. [32]
however, has been utilized in a similar form for enthalpy probes for measurements in
29
enthalpy probe to measure the temperature and enthalpy of microwave induced argon
plasma. A water cooled stainless steel probe was used to draw a plasma sample through
the inner diameter. The water temperature was measured at the inlet and exit planes. The
temperature rise in the water is due to the plasma flow. In order to calculate the plasma
temperature, tare data was collected with the probe in the plasma stream and no flow
through the probe. The change in temperature between the tare and experiment conditions
was used to calculate the enthalpy of the plasma from a simple enthalpy balance. The
study achieved positive results; however, the accuracy of the enthalpy measurement was
measurements of the same flow field. A stainless steel probe was used to measure the
introduced into the flow and quickly retracted to prevent melting. The study showed the
intrusive measurement does have an impact on the flow field. However, CARS
measurements were not taken simultaneously with the probe, so the cause of the
characterize the presence of the probe on the flame.[17] Tunable Diode Laser Absorption
Spectroscopy (TdLAS) was used to measure the path averaged temperature over a
Hencken burner flame.[17] Measurements were taken with the probe at two different
heights above the flame front, both with and without probe sample being collected.
30
CHAPTER 3
The first iteration of high pressure WSR operation was designed using a metal
reactor. The reactor retained the toroidal shape of the atmospheric set-up, [36] but it was
made from Inconel 625 with an alumina thermal barrier coating. While the thermal
properties of ceramics make them suitable for a combustion environment, the low
thermal conductivity and tensile strength make them prone to cracking. Atmospheric
operation demonstrated that the life of a ceramic reactor was very short, with most only
cracking. For experiments at atmospheric pressures, cracks in the reactor did not present
detrimental effects. There was minimal force on the reactor to open the cracks to the
extent that the combustion gases would communicate with the interior of the test rig.
When experimenting at high pressure, however, the pressure difference could cause the
cracks to propagate through the entire thickness of the reactor and allow combustion
products to escape the toroid. The use of a metal reactor would prevent cracking and
provide a positive seal of the combustion cavity, while a ceramic thermal barrier coating
would allow protecting the metal from temperature in excess of its melting temperature.
31
Due to the high thermal conductivity and relatively low melting temperature of
metals the reactor, however, required active cooling. The upstream half of the toroid is
water cooled. A clover leaf tube is set into grooves on the upstream instrumentation plate.
Water is fed to the tube, maintaining the cooling plate at a near constant temperature. The
Water Cooled
Upstream Side
Air
Air Cooled
Cooled
Exhaust
Exhaust Side
Side
The exhaust side of the reactor sits on a flange built into the exhaust housing. Air
passages are machined into the plate allowing air flow through the majority of the area.
Conduction from the hot reactor half to the air cooled flange provides cooling. The
exhaust section mates with an instrumentation flange to allow the reactor assembly to be
The upstream water cooled plate has four holes taped for instrumentation. The
Compression fittings seal each of the instrumentation holes. In addition to the reactor
32
instrumentation, the exterior metal temperature was measured by type-K thermocouples.
The steady state temperature of the reactor indicates if the cooling set-up is sufficient for
Since the use of a metal reactor was a departure from the set-up established during
concept. All of operating points during the initial experimental phase with the metal
reactor were at atmospheric pressure. The cooling was seen to be sufficient to maintain a
safe metal temperature at the lean conditions seen near blowout. As the reactor
equivalence ratio was increased beyond 0.5 the cooling struggled to keep up with the heat
release. Initial lean blowout conditions taken in the reactor showed an elevated blowout
equivalence ratio for the same flow rates, as compared to previous data [31] with a ceramic
reactor. The increase in blowout equivalence ratio is due to the heat loss from the reactor.
The cooling that is necessary for reactor health makes the reactor less adiabatic than
The HPWSR designed was modeled after the successful atmospheric WSR
design. The same 250 mL toroidal reactor section was used.[36] The toroidal section
promotes consistent mixing throughout the reactor volume. Previous experience showed
that while a metal reactor provides a consistent, positive seal, the combination of high
thermal conductivity and the need for cooling results in skewed blowout results. The heat
removed from the reactor to maintain structural integrity elevates the blowout
equivalence ratio. Therefore, a ceramic reactor was selected to achieve the most adiabatic
33
blowout data possible. The first design challenge arising from the use of a ceramic
reactor is the issue of cracking. As previously discussed, the low thermal conductivity
and tensile strength of the ceramic make it prone to cracking. When operating at
atmospheric pressure, the reactor operates at less than 1 psig over the ambient pressure.
With a very slight overpressure, leakage from the reactor due to cracks is minimal. When
adapting the hardware to operate at high or low pressure, however, the potential leakage
from even a small crack is dramatically increased due to the higher pressure differential.
Providing a positive seal to the reactor is critical to successful operation. Two sealing
rings were designed to address the problem of reactor containment. The two sealing rings
are machined to accommodate gaskets as well as seat the reactor. Mica gaskets with a
stainless steel core were used for testing. The sealing ring configuration is shown in
Figure 3.2.
Sealing Rings
Figure 3.2 shows a cross-section of the HPWSR assembly. The toroidal reactor
section is shown in purple, and is contained by the upstream water cooled plate, yellow,
34
and the air cooled exhaust housing, brown. Clamping studs, blue, provide the force to
seal the gaskets. The sealing rings are shown in light yellow and sandwich the jet ring.
The sealing rings provide a positive seal outside of the reactor and do not depend upon
In order to achieve high pressure operation, the reactor assembly from Figure 3.1
must be fit into an appropriate flange to interface with a pressure vessel. It would be
possible to run at elevated pressure without the need for a pressure vessel, but the
pressure differential across the reactor, and the possible leak paths are comparatively
greater. Enclosing the reactor in a pressure vessel allows the surrounding pressure to be
pressure vessel allows the reaction pressure to be precisely controlled by a back pressure
construction of the well stirred reactor housing, as the vessel environment can be
All instrumentation lines as well as combustion air, fuel and cooling flows were
brought into the pressure vessel to supply the reactor. Previous combustion rigs used to
mount the hardware in a flange with threaded holes to allow pass through the required
plumbing while fittings provided positive seal. This “instrumentation flange” approach
was used for the HPWSR, and allows interface with existing facilities. The supply
plumbing was designed to fit within the 13.5-inch diameter vessel that would be used for
testing. Figure 3.3 shows the HPWSR assembly inside the pressure vessel.
35
Pressure Vessel
Air
Manifold
Reactor
Assembly
In Figure 3.3, the instrumentation flange is shown in orange. All of the plumbing
must pass through the holes on the flange, and make a 90° turn upstream to reach the
reactor. Custom made fittings are used to seal tubing and instrumentation passing through
the flange. Air is fed through the instrumentation flange at two locations, 180° out from
each other. The air flows into an arc-section manifold upstream of the reactor. The
manifolds distribute the air to two feed tubes, which flow to the jet ring. Two manifolds
feed four inlet tubes. Previous WSR designs used either a single or double feed to the jet
ring. Less feed tubes tend to prioritize the inlet region for flow, and starve the region on
the opposite side. In order to promote even and consistent flow and mixing, four inlet
tubes were used. The inlet tubes were welded to the jet ring to eliminate compression
fittings, a possible source of leaks. A detailed view of the inlet tubes entering the jet ring
36
Figure 3.4: Jet ring inlet design
As shown in Figure 3.4, caps are welded on the ends of the tubes and a 0.062-inch
hole is drilled in the center. Additionally, two slots are cut in the side wall of the tubes.
The cap and slot design helps to promote bulk flow circumferentially around the jet ring,
while the hole in the cap prevents the jet directly under the inlet from being starved of
flow.
absence of an external ignition source. In order for a fuel to auto-ignite, it must be mixed
with an oxidizer within the flammability limits of the fuel. A typical hydro-carbon fuel
cannot auto-ignite in the absence of oxygen. Auto-ignition events are caused when a fuel-
air mixture is heated to a high temperature. For example, the auto-ignition temperature of
Jet-A is 410 °F while the auto-ignition temperature of hydrogen is 997 °F. [21] The time in
which the auto-ignition reaction requires to begin is called the ignition delay time, and is
37
a function of temperature, pressure, fuel used, and equivalence ratio. When operating in
premixed systems, such as the WSR, it is possible for auto-ignition to occur at any point
after the fuel and air are mixed given the proper conditions. Therefore, it is important to
A study by Lefevbre et al[22] showed that reaction inlet temperature had the
strongest influence on ignition time delay in lean propane-air mixtures. Lefevbre also
showed that the effect of varying equivalence ratio on ignition delay was weaker than the
effect of temperature. At a given fuel-air ratio, less ignition energy is required to initiate
The autoignition characteristics of fuels are important in gas turbine and reactor
design as the ignition delay and autoignition temperature represent the upper limit of
residence time and fuel-air mixture temperature that is acceptable in the system [22]. The
effects of ignition delay are more pronounced in applications where the primary heat sink
for cooled components is the fuel. In systems that utilize premixed combustion, such as
the WSR, short ignition delay presents a problem for large reactor volumes or long
residence times in the feed lines. This is because the reactants can auto-ignite in the feed
lines before entering the combustor. The issue is magnified with the use of heavy
hydrocarbons, such as JP-8, that require higher temperature to maintain the vapor state.
[21]
38
3.2.2. Fuel Vaporization
The atmospheric WSR used an external vaporizer is shown in Figure 3.5. The
large, heated volume of the vaporizer provides a long residence time for mixing to occur.
Fuel vaporization is achieved with a Delavan air assist nozzle, shown in Figure
The air assist nozzle shown in Figure 3.5 uses an air stream and swirler to break
up the fuel flow into small droplets. Smaller droplets evaporate more rapidly, and,
therefore, it is advantageous to obtain the smallest droplet size possible. When evaluating
the evaporation of droplets in a hot flow, the evaporation constant [21] is used. The
evaporation constant, as proposed by Lefevbre, takes into account the fact that the heat
the droplet absorbs during the heat-up period reduces the available heat for the
vaporization of the droplet, lowering the evaporation rate over a steady state
during phase change lowers the overall temperature of the environment, reducing the
energy available for evaporation of additional droplets. The effect of droplet cooling will
lead to longer evaporation times then when a steady state approximation is used.
Lefevbre defines the evaporation constant as . [21] Figure 6 shows the effect of
39
droplet size on evaporation time. Multiple evaporation constants are also represented to
1000
800
700
Evaporation Time (ms)
600
500
400
300
200
100
0
0 20 40 60 80 100 120
Droplet Size (um)
In Figure 3.6, the evaporation time as a function of droplet size is shown for three
different evaporation constants. The clear trend is that as droplet size increases, the
The internal passages of the nozzle are sized such that there is a small pressure
drop from inlet to exit, allowing the pressure to be driven by downstream conditions. The
coupling of upstream and downstream conditions was mitigated by the use of a needle
valve immediately upstream of the nozzle to produce a large pressure drop in the system.
The air flow rate through the nozzle was maintained constant over the majority of testing
40
The external vaporization technique, detailed above, has been used extensively
during atmospheric testing, and has been used for over several hundred hours of stable
operation. The main disadvantage of this vaporization scheme is the long residence time.
As the reactor pressure increases, the temperature required to maintain a fuel vapor also
increases. Increasing mixture temperature brings the mixture closer to autoignition. Long
vaporizer residence times further increase the risk of an autoignition event when
operating at high pressure. Therefore, the ability of the vaporizer to safely contain
The vaporization of JP8 poses a greater challenge than other lighter liquid fuels
since more heat is required to completely vaporize JP8, coupled with a lower autoignition
temperature. The current vaporizer design has been successfully and safely used to
vaporize a variety of liquid hydrocarbon fuels, including JP8, during atmospheric testing.
The WSR system is designed such that the pressure drop across the nozzles in the jet ring
is the largest in the system. Having the largest pressure drop located at the nozzles means
that as the reactor pressure is increased, the vaporizer pressure must be increased by a
larger factor. As the vaporizer pressure is increased, higher fuel temperatures are required
the dewpoint. The dependence of pressure on the dewpoint of JP-8 was presented by
Stouffer et al.[36] and is shown in Figure 3.7. A JP-8 surrogate was developed by Stouffer
41
of hydrocarbon mixtures. It is clear that as the pressure increases, more energy is required
to vaporize the fuel. Additionally, as equivalence ratio is increased, more fuel in present
in the mixture, driving the dewpoint even higher. The dewpoint dependence on
equivalence ratio becomes significant when investigating soot formation and rich-
500
450
Dewpoint Temperature (°F)
400
350
300
vaporizer will be so great that the temperature required to vaporize the fuel exceeds the
autoignition temperature. The autoignition temperature of JP-8 is 460 °F. In order to use
the existing vaporizer for high pressure testing, the test matrix will be limited such that
42
After analyzing the limitations of the vaporizer hardware available for the
experiments at high pressure, it was determined to limit the reactor pressure to 3 atm.
Based on this desired maximum operating pressure, the flow rates (calculated from
desired residence time and reactor volume), and the jet ring nozzle geometry, the required
upstream pressure could be calculated. It was found that the highest vaporizer pressure
that would be experienced during testing at 44 psia reactor operating pressure was 165
psia, corresponding to τ = 3 ms and ϕ = 0.3. Knowing the dewpoint for all of the test
conditions and the maximum pressure that would be experienced in the vaporizer, fuel
and air temperatures could be determined such that the mixture was a vapor at all
Figure 3.8 shows the variation of the mixture temperature with equivalence ratio, for an
for all operating conditions is desirable for safe operation. The next step in the analysis
was to ensure that the construction of the vapoirzer would stand up to a un-planned
that the pressure wave propagates at the Chapman-Jouguet speed, and thus passes
example of a Rankine-Hugoniot plot with the important plot features noted. [20]
The calculation of the detonation pressure rise requires that the ratio of specific
heats (γ), the reactant pressure and temperature, and the heat release. The ratio of specific
heats for JP8- air mixtures was given through SUPERTRAPP, while the reaction
temperature and pressure was defined through by the desired test conditions. The heat
44
release was determined by stoichiometry knowing the approximate composition of the
JP8 surrogate used in the analysis and assuming complete combustion. Substituting in the
values relevant to the desired test conditions yielded a maximum detonation pressure rise
of 10.14. A detonation pressure rise, P2/P1, of ~10 for stoichiometric JP8-air mixtures is
confirmed in the literature. [20] Based on the maximum pressure rise that could be
experienced during testing and the maximum vaporizer pressure (i.e.,165 psia) the peak
pressure to be taken by the hardware in the event of a detonation is 1650 psia. After the
peak pressure during a detonation was determined, the stresses on the vaporizer could be
calculated. The vaporizer structural analysis presented in Appendix B shows that the
When using an external vaporizer, the residence time of the mixture in the
plumbing is long, increasing both the heat loss from the fluid and the time the mixture
has to react. The high temperature needed to achieve the flash vaporization, required by
the internal vaporization technique, increases the autoignition risk. The system must be
capable of vaporizing and thoroughly mixing the fuel and air without burning upstream
of the jet ring. In order for the HPWSR to achieve the elevated pressures, a different
time and allows high reactor pressure. Additionally, the implementation of this system for
45
3.3.1. Experiments with a WSR at Elevated Pressure
vaporizer. The safety analysis in Section 3.2.3 showed that the vaporizer was structurally
capable of limited high pressure operation. Above 3 atm, however, structural limitation
mandated a different method. The external vaporizer was removed and the fuel lines were
routed directly into the pressure vessel to feed injectors mounted in the WSR air feed
lines. In order to achieve vaporization inside the pressure vessel near the jet ring, the
properties of a super critical fluid were utilized. A supercritical fluid is one where the
temperature and pressure are maintained above the critical point. In this state, there is no
distinct liquid or vapor phase. Additionally, relatively small changes in pressure result in
large changes in density. The internal vaporization technique uses orifice tubes that act as
throttling valves. As the high pressure fuel flows through the orifice into a significantly
larger area, the pressure drop causes flash vaporization. The orientation of the orifice
46
Secondary Fuel
Leg
Orifice
Tubes
Fuel Supply
Air Feeds
Figure 3.10 shows four of the orifice tubes used for high pressure testing. The
tubes (shown in red) are fed supercritical liquid JP-8, and spray directly into the 4 air-
feed lines (shown in green) leading to the jet ring. The air lines act as the mixing volume.
Utilizing flash vaporization allows the nozzles to be placed closer to the reactor, reducing
the residence time required for mixing. Smaller residence time allows for mixture
temperature potentially to reach levels that would auto-ignite in the higher volume,
The orifice tubes are closed end stainless steel tubes with a laser drilled orifice in
the end. The main disadvantage of using a fixed geometry nozzle is the limitation on
operating flow rate. Nozzle pressure drop is a function of flow rate, and a higher flow rate
47
will require a large pressure drop across the nozzle. The supercritical fuel must be
maintained at pressure, while the pumps are only capable of operating up to 3000 psi.
These limits define the orifice size that must be used. In order to increase the operating
range of internal vaporization two fuel flow circuits were installed utilizing 50 and 100
micron orifice tubes. During low flow situations, the 50 micron circuit is used. As the
flow rate in increased, the 50 micron circuit is turned off, and the 100 micron turned on.
If it is desired to run the flow rate higher both can be operated for a total of 150 micron
orifice size.
The final aspect of the internal vaporization design was a fuel line heat trace. The
fuel is heated to 650 °F, which is above the critical temperature. At these high
temperatures, the heat loss rate from the fuel between the heater and nozzles can be
significant. If the fuel falls below the critical temperature, it will pass through a two
phase region rather than moving directly to the vapor phase, which may result in
unsteady combustion. A 0.375-inch tube was fit around the fuel lines, with hot nitrogen
flowing through. The nitrogen was heated to 800 °F to allow some heat loss from the
flow without allowing the inner co-flow fluid to drop below the fuel temperature.
addressed, the final reactor assembly was installed in the pressure vessel. Figure 1 shows
the HPWSR installed in the instrumentation flange with all of the necessary plumbing.
The air manifolds, reactor stack-up and fuel lines can be seen. Additionally,
48
thermocouple and pressure instrumentation is passed through fittings in the flange
49
In Figure 3.11, flows proceed to the right, and the exhaust section can be seen in the
back on the right side of the figure. The facility controls reactor pressure with a back-
pressure valve. The back-pressure valve used for testing has a maximum temperature
rating. Air is used to quench the exhaust stream and prevent damage to the back-pressure
valve. The 1-inch lines that carry the exhaust quench flow can be seen at the midpoint of
the flange, on the right side in Figure 3.11. For the low air flow rates at which the WSR
operates, air quenching is sufficient to drop the exhaust temperature. For higher flow rate
vaporization technique
dilution lean blowout data and to demonstrate the extended capabilities of the
WSR
The following sections details the experimental procedure and test conditions used in
Prior to testing, all the air heaters and heated lines feeding the reactor are turned
on and airflow established. The pre-heating process serves two purposes. First, the hot
50
air heats all the metal plumbing that the fuel/air mixture will contact and heats the
ceramic in the reactor. When using a liquid fuel such as JP-8, all of the plumbing must be
maintained above the dewpoint of the mixture. If there are any cold spots in the system,
fuel vapor may condense and form liquid fuel droplets in the feed lines. Preheating the
ceramic reactor makes ignition easier and helps to reduce the thermal shock experienced
by the ceramic. During atmospheric testing, the liquid fuel is heated to 300 °F by a
Mokon oil heater. The oil is also used as a coolant for the emissions probe.
After the hardware preheating is complete, the air flow rate is decreased to 100-
150 SLPM. The igniter is activated and gaseous fuel incrementally introduced until light-
off is achieved. Successful ignition is verified with the type-B thermocouple in the
reactor volume. Ethylene is used as the fuel for light-off due to its wide flammability
equivalence ratio than JP-8 at the same conditions. During the light-off procedure, the
equivalence ratio is maintained below 0.6 to avoid unnecessary heat stress on the
ceramic. As the reactor temperature increases, air is added to maintain the temperature
below the melting point of the ceramic. The next step is to switch to liquid fuel. The
liquid fuel flow rate is slowly increased while simultaneously decreasing the gaseous fuel
flow rate. Changing both flow rates will maintain an approximately constant reactor
temperature. After the reactor is running entirely on liquid fuel, it is allowed to reach a
steady state temperature. Temperature traces in the data acquisition software are used to
indicate when the reactor temperature has stopped changing and reached equilibrium with
the fluid temperature in the reactor. The fuel and air flow rates are adjusted to the first
51
The air flow rate is metered using a thermal mass flow controller manufactured by
Brooks. Blowout is achieved by reducing the fuel flow rate while maintaining a constant
air flow rate. The timing of the fuel flow rate reductions is very important in the accuracy
of lean blowout data. If the fuel flow rate is decreased too fast the reactor will not be able
to equilibrate, and the walls will be at a higher temperature for a given equivalence ratio.
Hotter walls will tend to stabilize the flame, resulting in a blowout equivalence ratio that
is lower than the true adiabatic blowout. The goal of the blowout study is to begin to
determine the lean stability margins for JP-8 in a WSR. Several different air flow rates
were tested. The air flow rate changes the residence time in the reactor, and an increase in
the former results in an increased air loading parameter. Table 3.1 shows the test matrix
used for the atmospheric portion of this study. Each of the blowout conditions was
Air Flow
Fuel LP Starting
Rate (SLPM)
275 1.44
300 1.41
JP-8 0.6
500 2.23
650 2.84
The properties of new fuels cannot be guaranteed to be the same as those of JP-8.
Atmospheric pressure testing was conducted to evaluate the impact of fuel Cetane
ignition delay of certain fuel. DCN is typically applied to diesel and heavy aviation field,
52
such a JP-8 and its alternative variants. High DCN fuels have shortened ignition delay
times, and therefore are easier to ignite. A lower DCN results in a fuel that requires more
energy to ignite. In this manner, DCN is the opposite of the octane rating used for
automotive gasoline.
While the majority of properties of JP-8 are specified, there is no specification for
the DCN. The same is true for the alternative fuels that are gaining attention. If a specific
alternative fuel is shown to meet the density, flashpoint and lubricity requirements
similar DCN. The possibility of a widely varying DCN could mean that an alternative
address the impact of DCN on combustion, alternative fuels were selected that bracketed
the DCN of JP-8, one higher and one lower. The fuels selected for the study are shown in
Table 3.2. The goal of the study is to quantify what impact, if any, DCN has on the
C11.4H22.3
C10.8H23.2
C11.9H25.6
53
Table 3.2 shows the properties of the two alternative fuels, along with the
properties of JP-8 for comparison. The alternative fuels are iso-paraffinic kerosene (IPK)
manufactured by Sasol, and Hydro-Treated Renewable Jet fuel R-8 (HRJ). The average
fuel composition from GCxGC analysis is shown, along with the DCN.
Test conditions were chosen to achieve the largest loading parameter range
possible with the atmospheric reactor. The low flow rate was chosen based on initial
exploratory work that showed inconsistent blowout data with flow rates below 275
SLPM. The high end of the matrix was facility-limited to 675 SLPM. Three blowout
points were taken at each condition to ensure the results are statistically significant.
Air Flow
Fuel DCN LP Starting
Rate (SLPM)
275 1.44
325 1.71
375 1.97
JP-8 47.7 0.6
500 2.62
600 3.13
675 3.51
275 1.44
325 1.71
375 1.97
IPK 31.5 0.6
500 2.62
600 3.13
675 3.51
275 1.44
325 1.71
375 1.97
HRJ R-8 59.1 0.6
500 2.62
600 3.13
675 3.51
54
Lean blowout experiments were conducted using the metal reactor assembly. The
metal reactor assembly was installed in the Atmospheric Combustion Research Facility,
and, therefore, it uses all of the same equipment as the previous atmospheric WSR tests.
The procedure for warm-up and reactor ignition is also the same as in previous testing.
For each condition, the equivalence ratio was set at 0.6 and the reactor allowed to reach a
steady temperature. A constant measured temperature in the reactor indicated that the
ceramic walls were near the same temperature as the fluid, and heat loss was in
equilibrium with the heat release of the combustion. Fuel flow rate was then reduced,
Nitrogen dilution data was collected during the atmospheric test campaign. As
previously discussed, the excess nitrogen in the air will lower the combustion
pressure through kinetic theory. From the definition of loading parameter, there is an
pressure will yield large increases in air loading parameter, extending the stability curve
for the reactor. The literature [10] shows that nitrogen dilution can increase the air loading
parameter over 2 orders of magnitude without the need for pressure vessels or new
hardware. In order for the technique to be validated, nitrogen dilution data must be
The HPWSR hardware, with fused silica reactor was used to collect nitrogen
dilution data. The hardware was installed in the Atmospheric Combustion Test Facility.
55
The facility hardware and instrumentation is the same as with previous atmospheric
testing. Nitrogen was added to the air flow upstream of the heater. Mixing occurred over
Two different nitrogen mixing schemes were used in testing. The first was
nitrogen addition. When testing with nitrogen addition, the air flow rate was set to obtain
the desired residence time, and then dilution nitrogen was added. This method increased
the overall flow rate of the mixture into the reactor. The second method used was
nitrogen replacement. When using the nitrogen replacement method, the air flow rate was
decreased as nitrogen flow rate was increased to maintain the same combined overall
mass flow rate of nitrogen and air. The mass flow rate calculation accounted for the
difference in molecular weight between air and nitrogen. Table 3.4 shows the test matrix
56
Table 3.4: Conditions for Nitrogen Dilution Testing
Nitrogen
Air Flow
Test Flowrate
Rate (SLPM)
(SLPM)
500 5
500 25
Nitrogen 500 35
Addition 500 50
500 75
500 100
475 25
450 50
400 100
350 150
Nitrogen
275 267
Replacement
275 225
262 238
258 242
254 246
The conditions shown in Table 3.4 represent both the nitrogen addition and
nitrogen replacement. Higher blowout equivalence ratios and, therefore, lower simulated
Fuel was heated with a 3kW sand bath for all conditions. The sand bath is
fluidized by nitrogen. The fluidization process greatly increases the heat transfer between
the medium and the fuel, while still allowing high temperatures to be reached. The heat
transfer to the fuel was further maximized by using a 20-foot coil of 1/8 inch tubing. The
small diameter of the tubing resulted in high velocity, while the length provided a large
surface area over which heat transfer would occur. Prior to heating, the fuel was sparged
57
with nitrogen for a day. The bubbling of nitrogen helps to remove dissolved oxygen from
the fuel, helping to prevent oxidative coking that may clog the fuel lines at high
temperature. When testing with the external vaporizer, the fuel was heated to 350 °F.
The fuel was delivered with ISCO 1000D continuous flow syringe pumps. The
ISCO pump system is capable of delivering 2500 psi fuel pressure, and delivering a flow
rate of .01-204 ml/min. The pumps meter the flow, so no additional flow meters were
the main LabView program. This also permits the fuel flow rate and pressure data to be
The combustion air for experiments at elevated pressure was heated with a 15kW
circulation heater, while the combustion air for early atmospheric experiments was heated
with a 3 kW circulation heater. Flow was provided using the shop air at the facility and
had a maximum pressure of 100 psi. The air flow was metered using a CFM025 Coriolis
meter. The Coriolis meter has a large turndown ratio, thus allowing the same flow meter
to be used over a large range of test conditions. The general flow path used for testing
58
Figure 3.12: Flow path schematic for experiments with an external vaporizer
One of the main concerns when using a Coriolis flow meter is the back pressure.
Due to the construction of the meter, a large pressure drop can cause the internal passages
to choke, potentially damaging the meter. In isentropic air, the pressure drop required to
choke the fluid is 52.5%. For the safety of the hardware, the maximum allowable
pressure drop across the Coriolis meter was set to 25%. The back pressure and overall air
flow rate was regulated with two pressure operated, fail closed, Badger Research control
valves. Two different valve trims, with different valve flow coefficients (Cv), were used
to allow fine and coarse flow control. The coarse control valve had a Cv of 2, as was used
to set the flow rate and pressure in the general area desired for the specific test condition.
The coarse valve was manually set to the same setting until a large change in the air flow
rate was needed. The fine control valve had a Cv of 0.5. This valve was controlled with a
PID loop within a controller. After the desired flow and meter backpressure was dialed in
59
the fine control valve was set to automatic mode, and the flow rate was maintained for
the test.
3.6.1. Sub-atmospheric
external vaporizer, the same setup was used for low pressure testing. The main
consideration in adapting the atmospheric vaporizer hardware to low pressure testing was
to manage heat loss. In order to maintain a constant reactor residence time, as the reactor
pressure decreases, the air flow rate must be decreased as well. The low flow allows more
time for the fluid to lose heat through the plumbing. A heat trace was applied to the fuel
tubing to minimize heat loss. The fuel line was passed through a 0.375-inch line, which
was attached to a 1200 W heater. Nitrogen was heated and flowed along the fuel lines.
Extra ceramic fabric insulation was added on the exterior of the plumbing throughout the
entire system to further combat heat loss from both the air and fuel. Additionally, tape
heaters were wrapped around the vaporizer body, the combustion air inlet and the
pressure vessel. These heaters were controlled to 300 °F and served as system preheat.
The final aspect of thermal management was heated nitrogen that was flowed into the
pressure vessel. The nitrogen was heated to 350 °F and served three purposes. The first
was to preheat the reactor assembly and internal supply lines, while the second purpose
of the nitrogen was to provide an inert environment inside the pressure vessel as an extra
safe guard against leaking fuel/air mixture. The final purpose of the nitrogen was that the
pressure difference between the reactor and the vessel could be minimized, theoretically
60
Blowouts were achieved at three (3) different sub-atmospheric pressures. The
lowest pressure the facility was capable of achieving was 4 psia. In order to further
increase the loading parameter the air flow rate was increased at the lowest pressure. The
The first set of high pressure lean blowout data was conducted with the external
vaporizer, allowing more direct comparison between atmospheric and low pressure
blowout data taken with similar hardware. As discussed in Section 3.3.1 the external
vaporizer could only safely support testing up to 3 atm. The initial testing, then, was used
to validate the hardware and reactor stack-up for its suitability at high pressures. The
ability of the reactor to maintain a positive seal, the ability of the cooling flows to
maintain acceptable operating temperatures and the controllability of the air and fuel
flows at the higher pressures were of specific interest. The same flow set-up was used for
the high pressure testing as was used for the low pressure testing. The 15kW heater was
sufficient to provide the required air temperatures for high pressure operation. The test
conditions used for reactor checkout and initial high pressure blowout data are shown in
Table 3.6.
61
Table 3.6: Test Conditions for High Pressure Blowout with External Vaporization
Air Flow Reactor
Fuel Rate LP Pressure Starting
(lbm/min) (psia)
2.84 3.25 29.45
JP-8 4.24 3.0 38.18 0.6
3.90 2.0 45.50
TECHNIQUE
While high pressure testing with the external vaporizer served as a proof of
concept and afforded time to validate the hardware and operating procedure, the safety
limitation of the vaporizer prevented the reactor pressures from reaching the target of 5
atm. The next phase of testing was to evaluate the internal vaporization technique, and
utilize the lower residence time to achieve ever higher reactor operating pressures.
With the internal vaporization technique, the fuel is heated with the fluidized sand
bath to 750 °F in order to facilitate supercritical flash vaporization. For the high pressure
testing, 100 and 50 micron orifice tubes were installed on the two fuel legs. This
combination of orifices provide adequate pressure drop to achieve vaporization while still
The temperature of the nitrogen heat trace over the lines was increased to 800 °F.
Now that there was no longer a vaporizer body and fuel nozzle, the air from the 15kW
heater was routed directly into the air feed lines. The same thermal management
techniques (i.e., tape heaters and ceramic insulation) used in the external vaporization
to demonstrate the feasibility of the direct vaporization method and to establish some
baseline operability data for the reactor. The second is to use the new technique to
achieve blowout data at a reactor pressure of 5 atm. Additionally, several blowout points
previously tested with the external vaporizer were tested to evaluate the repeatability of
the new vaporization technique. The test conditions used are shown in Table 3.7Table
3.7.
Table 3.7: Test Conditions for High Pressure Blowout with Internal Vaporization
3.7.2. Sub-atmospheric
After the internal vaporization technique was tested at high pressure, it was
decided to evaluate the performance of the system at low pressure. The low pressure
vaporization is the difficult to maintain due to the high heat loss associated with low air
and fuel flow rates. The negative impact of heat loss is partially offset by the reduced
dewpoints caused by sub atmospheric pressures. If any of the orifice tubes is not
producing complete vaporization, but rather producing fuel droplets, the impact would be
more evident in the reactor at low pressure. The 50 micron orifice tubes were used, while
all of the other systems remained the same. Several of the low pressure conditions tested
with the external vaporizer were repeated with the internal vaporizer. Additionally, a final
63
high flow rate, low pressure condition was conducted to maximize the loading parameter.
The design of the temperature probe is discussed in Section 2.5.2. In practice, the
choked conditions required to measure the mass flow are achieved with orifices. The
ceramic probe is constructed from a 0.375-inch diameter extruded alumina tube with a
closed end. A 0.062-inch hole is drilled into the side of the tube with a diamond-burr bit.
The orifice in the side of the probe creates a hole in cross-flow condition. The orifice in
the ceramic is inserted into the flame. Downstream of the probe, a 0.062-inch orifice
fitting is inserted into the system. The downstream orifice meters the system and
provided the temperature and pressure measurements required for the temperature
calculation.
The ceramic portion is inserted into the flame, where the pressure is known.
Sample is drawn through the system to a second orifice, there the temperature and
pressure are measured. The sample is collected using a Venturi vacuum pump. The pump
64
is connected to 80 psi shop air, and it is capable of sustaining a vacuum down to 3 psia. A
low downstream pressure is essential to ensure that both of the orifices are choked.
Starting at atmospheric pressure, the first orifice must be bought down to 7.35 psia, while
the second orifice will need to be at 3.675 psia for the system to be choked, assuming
isentropic, ideal flow. The differential pressure across the second orifice is measured
thermocouple inserted downstream of the orifice fitting. Figure 3.13 shows the set-up that
In Figure 3.13, the pressure tap nearest to the vacuum pump was used to ensure
that the system was choked. The probe static pressure tap, indicated in the figure, is used
65
in the calculations, and is a function of the gas temperature. The thermocouple indicated
is a K-type and is used to measure the fluid temperature immediately after the second
When the system is used, the probe was set to the desired location within the
flame, and the vacuum pump activated. The vacuum pump was controlled by changing
the pressure of the shop air. After the vacuum was established, the system was allowed to
preheat. In most cases, it was not possible to choke the orifice system at ambient
temperature. The second orifice had to exceed 100 °F before choked flow could be
established in both orifices. After it was verified that the system was choked, the
The first test of the temperature probe was to compare the temperature calculated
with the probe with thermocouple readings taken in close proximity to the probe orifice.
and accuracy of the probe upon which to base further testing efforts. The probe was
evaluated over two different burners. The first was in the exhaust of the WSR. The
exhaust of the WSR provides a high flow rate and a high heat release environment that
the probe is unlikely to perturb. The second flame tested was the flat flame of the
McKenna burner. This burner has a smaller flame and has lower flow rate. However, the
flame is very steady, optically accessible, and well characterized. The McKenna burner
represents a case when the presence of the probe may affect the flame. The conditions
66
Table 3.9: Test Conditions for Ceramic Probe to Thermocouple Comparison
400 0.60
400 0.70
WSR
400 0.80
400 1.4
12 0.6
12 0.7
McKenna
12 0.8
Burner
12 0.9
12 1
The second phase of the probe characterization was to compare the results calculated
from probe measurements to temperatures taken with laser diagnostics over well
temperature profile of a Hencken burner flame with two different fuels. This study was
repeated and temperature measurements were taken with the ceramic probe. The
temperature results from CARS can be considered the true adiabatic temperatures of the
flame, and are considered the target for the calculated probe temperature. The test
67
Table 3.10: Test Conditions for Hencken Burner Test [32]
12.2 0.5
12.2 0.6
12.2 0.7
12.2 0.8
12.2 0.9
H2 12.2 1
12.2 1.1
12.2 1.2
12.2 1.3
12.2 1.4
12.2 1.5
12.2 0.51
12.2 0.62
12.2 0.71
12.2 0.81
12.2 0.91
C2H4 12.2 0.96
12.2 1.01
12.2 1.06
12.2 1.11
12.2 1.21
12.2 1.3
flame, TDLAS data was taken simultaneously with the probe to characterize the intrusive
impact of the probe in the flame. TDLAS data was taken in connection with another
research group. [17] Data was collected at two different heights above the burner. The
orientation of the probe is shown in Figure 3.14. At each height, data was collected both
while the probe was pulling sample and when the vacuum pump was off and no sample
68
was being collected. Baseline TDLAS temperatures were recorded without the probe in
the flame.
Data was collected using ethylene as the fuel. The conditions were a truncated
matrix of the conditions used in the CARS characterization and are shown in Table 3.11.
12.2 0.8
C2H4 12.2 1
12.2 1.2
The experiments conducted with the ceramic temperature probe showed that the
environments. However, due to a tight time constraints in the High Pressure Combustion
Facility, the probe was not used as part of the HPWSR campaign. The results from the
69
CHAPTER 4
The first test conducted with the new HPWSR hardware was a study of the
combustion stability of alternative fuels. The suitability of a cooled metal reactor for
blowout studies was also investigated. The specific focus was to investigate the impact of
cetane number of the blowout characteristics. The cetane number is an indicator of how
easily a fuel will ignite in a compression ignition environment, and is often used as a
measure of diesel fuel quality. Specifically, cetane number is the inverse of octane
number, and inversely proportionate to the fuels ignition delay. A fuel with a higher
cetane number will have a shorter ignition delay time. Currently in the JP-8
specifications, there is no target cetane number, or cetane number range, for new
alternative fuels. As alternative fuels are refined from varying feed stock, there is the
possibility for a wide variation in the cetane number of those fuels relative to JP-8. When
these new fuels are used in systems designed to run on JP-8, the change in the ignition
delay time has to potential to effect the ignition and stability limits of the combustor.
Initial studies in the WSR are intended to determine if there is a noticeable change in the
lean blowout characteristics of alternative fuels from the JP-8 baseline, and quantify the
strength of the correlation between lean blowout and cetane number. The specific test
70
conditions for the study are shown in Table 3.3. While it was intended to compare the
blowout of two fuels to JP-8, limited lab time only allowed one alternative fuel to
becharacterized. The fuel used was iso-paraffinic kerosene (IPK) with a cetane number of
31.5. The comparison between lean blowout of JP-8 and IPK is shown in Figure 4.1
The data shown in Figure 4.1 indicates that the equivalence ratio at blowout has a
weak dependence on fuel cetane number. A 34% increase in the fuel cetane number
caused only a 2.8% increase in the equivalence ratio at blowout. IPK has a lower cetane
number than JP-8, meaning that the ignition delay time of IPK is higher than JP-8. As a
result of the increased ignition delay time, IPK is more resistant to ignition and prone to
71
lean blowout at lower velocities of LP than JP-8. However, the dependence of lean
blowout on the cetane number was weak in the current study. Additionally, over the LP
range tested, the blowout equivalence ratio increased by approximately 10%, creating the
beginning of a stability curve. Previous work by Colket et al.[24 investigated the impact of
cetane number in a practical combustor. Seven fuels were tested with Cetane number
ranging from 30 to 80. The results from Colket et al. are shown in Figure 4.2. They
showed that the lean blowout equivalence ratio decreased the derived Cetane number
increased.
The results from Colket et al. [24] correspond with the results taken with the metal reactor.
72
In the practical combustor used, IPK produced a blowout equivalence ratio 3.3%
higher than Jet-A. This agrees well with the results from the metal reactor of a 2.8%
previous atmospheric data collected. Over a similar LP range with a ceramic reactor, little
variation in blowout is observed. The blowout data collected with the metal reactor is
biased to a higher equivalence ratio by the cooling. The difference in the LBO between
the fused silica ceramic and metal reactor for similar loading parameters is shown in
Figure 4.3.
operating temperature, increasing the heat loss from the reactor and causing an artificially
73
high blowout point. The ratio of the thermal conductivities of the two materials ( )
is 16. Therefore, for the high pressure test metal reactors are not desirable. However, the
increase in blowout caused by a change in cetane number is still applicable since both
After studies with the metal reactor showed that heat loss artificially increased the
the HPWSR can be found in Section 3.2. The following sections present data taken with
The first step to characterize the new HPWSR design was to repeat atmospheric
blowout data collected with the previous WSR design. This allowed a comparison of the
data to ensure that the HPWSR design was not artificially impacting the blowout data due
to high heat loss. The atmospheric data was collected over a LP range of 1.5-3, and is
74
Figure 4.4: Lean blowout in an atmospheric WSR
blowout equivalence ratio, over the LP range tested. The main limitation with
atmospheric experiments is the maximum mass flow rate attainable for the given reactor
volume. From the definition of loading parameter, the loading can be increased by
increasing the mass flow, lowering the pressure, or lowering the volume. In the initial
experiments, the pressure and volume were fixed, limiting achievable loading parameter.
Operating at low pressure increases the loading parameter, while operation at high
pressure allows more mass to be flowed for a given equivalence ratio. The limited
loading parameter range and minimal change in blowout equivalence ratio correspond
with previous blowout work with a ceramic WSR.[37] The results in Figure 4.5 indicate
that the new HPWSR design is performing as intended and is not skewing the data.
75
Therefore, the ceramic reactor and the HPWSR hardware were ready for testing at low
WPAFB.[6] The pressure in the reactor was varied from 4 psia to 73.5 psia. In addition to
described in Section 3.3.1, was evaluated. The lean blowout data as a function of LP is
76
Figure 4.6: Combustion stability curve for JP-8
approximately 28% was observed. The data show that the flame becomes less statically
significantly larger than the achievable LP range operating at atmospheric pressure. The
increase in the blowout equivalence ratio due to higher air flows and lower pressures (i.e.,
high LP) enabled the plotting of the lean side of the stability curve for a WSR during JP-
8.
77
Test data across multiple test days is also called out in Figure 4.6. The test data
recorded at near atmospheric pressures on over multiple days of testing correlate very
well. This consistency shows that the hardware is capable of generating repeatable data, a
achieved with both the internal and external vaporization techniques showed close
agreement. The comparison between the two vaporization methods is shown in Figure
4.7.
Figure 4.7: Blowout points taken with internal and external vaporization
78
A separate way of plotting combustor blowout data is equivalence ratio at
blowout vs. the combustor operating pressure. This type of combustor operating map is
useful in a gas turbine engine to demonstrate the operating pressures required for safe
operation. The lean blowout data as a function of reactor pressure is shown in Figure 4.8.
The data in Figure 4.8 clearly shows that as the reactor pressure is decreased, the
blowout equivalence ratio is increased, with the highest blowout point occurring at the
lowest reactor pressure or 4 psia. Low pressure operation represents the most highly
loaded condition and, therefore, illustrates fully the combustion limits of a given fuel.
Conversely, as the reactor pressure is increased, the flame becomes more stable and
blowout equivalence ratio is decreased, with the lowest blowout occurring at the highest
reactor pressure of 73 psia. As evident in the data, it is desirable to operate at the highest
79
pressure possible in order to achieve the most stable reaction. Figure also shows the
reactor operability map. Data sets taken over multiple test days at the same pressure have
very close correlation with each other. Additionally, data taken using external and
internal vaporization at the same condition agree closely, indicating that the fuel is
completely vaporizing and the internal vaporization technique is not biasing results. The
tight correlation between these two data sets shows that the internal vaporization
After reactor stability was established for low pressure conditions, the operability
was investigated for elevated pressures. Over a night of testing, the reactor pressure was
maintained at a constant 5 atm and the equivalence ratio was varied between 0.5 and
blowout. During this investigation, emissions data were collected. Figure 4.9 shows
80
When evaluating the emissions of gas turbine engines, it is important to vary the
conditions to discover the local minima, called the bucket. The emissions bucket is the
condition at which the emissions of a given species are the smallest. Emissions of CO
represent the inefficiency of the combustion reaction. If the combustion is more efficient,
more of the CO will be converted to the final product of CO2. As the equivalence ratio
approaches blowout, the combustion temperature decreases, and the conversion from CO
to CO2 is retarded. The equivalence ratio excursion in the WSR at high pressure was
sufficient to show the CO bucket for JP-8. Figure 4.9 shows that the reactor is operating
most efficiently, from CO to CO2 conversion, at an equivalence ratio of 0.46. When the
For the HPWSR at 5 atm, the minimum CO point occurred around an equivalence ratio of
0.46.
Emissions data were also collected at atmospheric pressure using the same
4.11 shows CO emissions data collected at 1 and 5 atm in the same reactor.
Figure 4.11 shows that significantly less CO was produced at 5 atm reactor
pressure than during atmospheric operation. Higher reactor pressures favor conversion
Additionally, the equivalence ratio at the minimum CO point is shifted lower at higher
82
pressures. The minimum CO point at 5 atm is at an equivalence ratio of 0.46, while the
These concentrations, on a volume basis, are shown in Figure 4.12. The emissions of both
CO2 and H2O increase linearly with equivalence ratio and have similar magnitudes.
The theory of low pressure simulation is discussed in Section 2.1.2. Diluting the
air with nitrogen allows a significantly higher loading parameter than otherwise
83
achievable in the same reactor. Data collected at atmospheric pressure using the nitrogen
Figure 4.13: JP-8 Combustion stability curve using simulated low pressure
curve for JP-8. Previous atmospheric tests showed minimal change in the blowout
equivalence ratio. The data in Figure is plotted against the nitrogen to fuel mass ratio.
During the nitrogen addition tests, excess nitrogen was added in addition to the
established air flow rate, effectively increasing residence time. In nitrogen replacement
tests, the overall mass flow rate to the reactor was maintained constant. As shown in the
achieved.
84
While the data collected has the appearance and overall trend of a JP-8 stability
curve, LP cannot be calculated until the nitrogen flow rate is correlated to a true low
pressure. Additionally, the peak blowout equivalence ratio in the study is 0.9. Theory
would suggest that the peak blowout equivalence ratio should be unity. The depression of
the peak blowout equivalence ratio could be a result of the incoming diluent causing
areas locally leaner than the bulk flow, creating combustion instability. Additionally, the
high concentrations of nitrogen at the large dilution ratios could be locally reducing
mixing, causing the reactor to blowout below unity. Further experiment and simulation is
A full calibration of the nitrogen dilution technique was not attempted in the
current study. To fully characterize the method, nitrogen dilution data would be
compared to true low pressure data of the same residence time in the same reactor.
However, the low pressure data points of the current study were sufficient to begin to
visualize a calibration. A comparison of low pressure and nitrogen dilution data is shown
in Figure 4.14, with the true low pressure data plotted on the upper x-axis, and nitrogen
dilution data plotted on the lower x-axis. The equivalence ratio at blowout is shown on
85
Figure 4.14: Correlation of true low pressure to nitrogen dilution data
Data collected at atmospheric pressure and 10 psia correlate very well with
nitrogen dilution data. Past 10 psia, however, the true low pressure data displays a
shallower stability curve slope than the nitrogen dilution data, resulting in nitrogen
dilution over predicting the blowout equivalence ratio at 4 psia. While the data set is not
sufficient to fully anchor the simulation, Figure 4.14 indicates that beyond a nitrogen to
fuel mass ratio of 6, the nitrogen in the system has additional impact of the blowout.
86
CHAPTER 5
CONCLUSION
The experimental study presented in this thesis detailed the development and
initial experimental entries using a WSR at high and low pressures. Previous work using
a WSR for gathering emissions data, fuel characterization and stability mapping had been
done at atmospheric pressure. In order to more accurately represent the conditions present
low pressures encountered at high flight altitudes put more stress on the stability of gas
turbine fuels. Therefore the ability of a WSR to operate at the two pressure extremes seen
in gas turbine aircraft allows the generation of benchmark quality data for emissions and
For the initial experimental studies with the high pressure WSR, fuel was
vaporized and mixed with air externally. The method of external vaporization has been
used successfully in the past for atmospheric WSR work. External vaporization was a
87
To push the WSR to higher pressures, a low residence time vaporization
technique was developed. Fuel was heated above its vaporization temperature, to a
supercritical state, and injected directly into the air lines upstream of the WSR jet ring.
Flash vaporization occurred through the use of orifice tubes, and mixing occurred in the
feed lines. The residence time was maintained shorter than the autoignition time for the
fuel so no burning would occur upstream of the reactor. This method proved stable and
increase in the loading parameter. The WSR hardware was also used to gather data at
sub-atmospheric pressure, and stable operation was demonstrated down to 4 psia. Low
pressure data applies to aircraft operating at high altitude, where combustion stability is
critical.
While gathering true low pressure data is important to fully understanding the
of this study was to determine if a correlation between low pressure combustion stability
and combustion stability with nitrogen dilution can be generated. Using nitrogen dilution
to simulate low pressures in an atmospheric rig can allow quick and simple evaluation of
5.2. CONCLUSIONS
The WSR was demonstrated to operate stably over a pressure range of 4 psia
to 73 psia.
88
The loading parameter achieved in the WSR was extended by a factor of 20,
40 .
engine.
Adiabatic conditions are required to meaningful blowout data. The low thermal
conductivity of fused silica makes it an ideal material for lean blowout studies.
Nitrogen dilution was shown to simulate the effects of low pressure over a
Cetane number was shown to have a weak impact of the blowout equivalence
89
A ceramic temperature probe using isentropic relations was shown to
radiation losses.
The ultimate goal of the high pressure WSR was to develop a complete stability
loop for JP-8 in conditions similar to a gas turbine engine. In order to continue to develop
the stability curve, higher pressures and mass flows are required. Further study with the
high pressure WSR should continue to adapt the hardware for operations at pressures up
to 20 atm. A 20 atm. capable WSR would allow lean blowout and pollutant emissions
data to be collected for JP-8 that is directly applicable to the conditions seen in gas
turbine engines.
The experiments with nitrogen dilution in the current study showed a promising
correlation with true low pressure data; however, time and facility availability constraints
prevented a full characterization. Future work should focus on taking nitrogen dilution
and low pressure blowout data in the same reactor. The direct comparison between
nitrogen and low pressure blowout data will allow a full correlation that can be used to
better understand future combustion systems and alternative fuels during a preliminary
design phase, without the need for the expensive low pressure combustion facilities.
the probe showed promising results on both laboratory flames and the WSR exhaust,
further design work is needed to optimize the size and material of the probe, the orifice
90
dimensions and the probe hardware and data collection system. Further, incorporation in
a full scale combustor test would demonstrate the probes ability to measure temperature
91
BIBLIOGRAPHY
Master's Thesis, Air Force Institute of Technology, Wright Patterson Air Force Base,
2011.
6. Dale T. Shouse, Jeff Stutrud, Charles Frayne. "High Pressure Combustion Research at
the Air Force Research Labortory." (International Society for Air Breathing Engines)
2001.
7. Edwards, Tim. "Advancements in Gas Turbine Fuels from 1943 to 2005." (Journal of
92
8. G. J. Sturgess, D. Shouse. "Lean Blowout Research in a Generic Gas Turbine
Combustor with High Optical Access." (International Gas Turbine and Aerospace
Astronautics) 1997.
Emissions from a Well-Stirred Reactor." (Journal of Propulsion and Power) 15, no. 2
(1999).
13. J. Zelina, D. R. Ballal. "Combustion and Emission Studies using a Well Stirred
15. James M. Gere, Barry J. Goodno. Mechanics of Materials. Cengage Learning, 2009.
16. Joseph Zelina, Richard C. Striebich, Dilip R. Ballal. "Pollutant Emissions Research
93
17. Justin T. Gross, Scott D. Stouffer, John Hoke, Drew Caswell, David L. Blunck, Fred
18. K. K. Gupta, A. Rehman, R.m. Sarviya. "Bio-Fuels for the Gas Turbine: A Review."
19. L. Rye, C. Wilson. "The Influence of Alternative Fuel Composition on Gas Turbine
20. Law, C. K. Combustion Physics. New York, NY: Cambridge University Press, 2006.
22. Lefevbre, A., Freeman, W., Cowell, L.,. ""Spontaneous Ignition Delay Characteristics
23. Mattingly, J.D. Elements of Gas Turbine Propulsion. New York: McGraw-Hill, Inc,
1996.
24. Med Colket, Steve Zeppieri, Zhongtao Dai, Don Hautman. "Fuel Research at UTRC."
25. Megan F. Karalus, K. Boyd Fackler, Igor V. Novosselov, John C. Kramlich, Philip C.
"Ignition and Emission Characteristics of Surrogate and Practical Jet Fuels." (Energy
94
27. Nenniger, J.E., Kridiotis, A., Chomiak, J., Longwell, J.P., Sarofim, A.F.,.
28. O'Neil, Alanna Rose. Chemiluminescence and High Speed Imaging of Reacting Film
2005.
34. Shazib A. Vijlee, John C. Kramlich, Ann M. Mescher, Scott D. Stouffer, Alanna R.
Toroidal Well Stirred Reactor." (Proceedings of the ASME Turbo Expo) 2013.
95
35. Simon Blakey, Lucas Rye, Christopher Willam Wilson. "Aviation Gas Turbine
36. Stouffer, S.D., Ballal, D.R., Zelina, J., Shouse, D.T., Hancock, R.D., Mongia, H.C.,.
37. Stouffer, S.D., Pawlik, R., Justinger, G., Heyne, J., Zelina, J., Ballal, D.,.
for Low Volatility Hydrocarbon Fuels"." Cincinnati, OH: 43rd Joint Propulsion
Conference, 2007.
38. Terrence R. Meyer, Sukesh Roy, Tohmas N. Anderson, Joseph D. Miller, Viswanath
39. Turns, S.R.,. An Introduction to Combustion. 2nd Ed. New York: McGraw Hill, Inc.,
1996.
41. Weiss, John P. Longwell and Malcolm A. "High Temperature Reaction Rates in
42. Zelina, J.,. ""Combustion Studies in a Well Stirred Reactor"." PhD Thesis, University
96
APPENDIX A
The first test with the temperature probe was a thermocouple comparison in the exhaust
of the WSR. The data for the thermocouple and the probe calculated temperature are
97
The calculated probe temperatures agree within 4% of measurements taken with a
thermocouple at the same height. When used to evaluate a high flow rate, high heat
release flame such as the WSR exhaust, the probe is very capable. The second
comparison for the probe was over a McKenna flat flame burner with an ethylene flame.
thermocouples, with a 9% error or less observed. Comparing the results of the WSR
exhaust and the McKenna burner, it is clear that the intrusive impact of the probe is
98
While thermocouple comparisons provided a quite diagnostic for probe
CARS data taken over the same flame. Data from Hancock et. al[32] from a CARS
used to compare the probe calculated temperatures. The data for the two fuels tested are
99
Figure A.4: Comparison of temperatures in an ethylene Hencken burner flame [32]
When testing over a small, low flow rate flame on the Hencken burner, the heat
loss due to the probe was evident. The calculated probe temperature was 12-18% lower
than the CARS measurements at the same location. The Hencken burner has a flame exit
area that is 1-inch square, while the outer diameter of the probe is 0.375-inch. The most
likely explanation for the low calculated temperature is the conduction losses form the
flame to the probe are so great that the probe temperature is not reaching the full flame
temperature.
The intrusive impact of the probe in the Hencken burner was characterized using
TdLAS.[17] Laser measurements were taken at two different heights over the burner, both
100
while sampling and not sampling. These measurements were compared to a baseline
without the probe present in the flame at all. The results are shown in Figure A.5.
The data show that at the lower height, while pulling sample, the flame
temperature, as measured with TdLAS is lower than the baseline, while at the same
height with no sampling results in a temperature higher than the baseline. At the second
height, the test when pulling sample correlated well with the baseline, while with no
sample the temperature was higher than the baseline, although by a lesser amount than at
height 1.
The discrepancies observed with the TdLAS survey can be explained by two
phenomena. Frist, when sample is being pulled through the probe, the movement induces
101
velocity in the surrounding flow. Since the orifice is choked, the flow approaches sonic
velocity as it nears the orifice. This velocity aerodynamically cools the flow, resulting in
additional cooling over the conduction losses, causing the temperature to be lower than
the baseline. When the sample is turned off, the hot ceramic insulates and stagnates the
heat from escaping from the top of the flame, causing an increase in the measured
temperature over the baseline. In the case of the Hencken burner, the probe diameter is
such a large percentage of the burner area that the insulation effect is noticeable. The
same principles apply at height 2, however, since the probe is more removed from the
CORRECTION
After characterizing the probe with laser measurements, it was decided that the
major factor in the temperature discrepancy was the losses due to the probe. The
Section 2.5.3. This correction was applied to data collected in the Hencken burner,
McKenna burner and the WSR exhaust. The results are shown in Figure A.6, Figure A.7,
102
Figure A.6: Comparison of corrected probe hydrogen flame temperatures
103
Figure A.7: Comparison of corrected probe ethylene flame temperatures
104
Figure A.8: WSR exhaust temperature with probe radiation-conduction correction
105
Figure A.9: McKenna burner ethylene flame temperature with radiation-
conduction correction
The correction improved the correlation to measured data across all conditions. The
Hencken burner with hydrogen flame was improved from 10-15% error to 0-6% error.
The Hencken burner with ethylene flame was improved from 13-18% error to 1-14%
error. Both of the thermocouple comparisons were slightly improved as well, from within
7% for the WSR to within 4%. The McKenna temperature was improved from within 9%
to within 4%. The correction produced a better correlation with the higher temperature
hydrogen flame. In the high temperature flame of hydrogen, the conduction and
106
APPENDIX B
calculated based on a peak loading of 1650 psia. The detonation pressure rise is discussed
in Section 3. The first stress calculated was the hoop stress. Classical thin-wall stress
analysis [15] was assumed for the hoop stress on the vaporizer, shown in Equation B.1
Equation (B.1)
The vaporizer is constructed from 304 stainless steel, which as 500 F has an
approximate yield strength of 24 ksi.[15] Knowing the value of the yield strength at the
elevated temperatures that will be present in the vaporizer, and the maximum
107
hoop stress that would be experienced, a factor of safety of 1.7 was calculated for the
current construction. This analysis shows that the vessel is capable of handling detonation
of JP8 under the worst conditions. Note that a pressure rise of 10.14 corresponds to
operation at ϕ = 1. This is a worst case scenario for the test matrix being presented, and
the majority of testing will occur well into the lean regime, below ϕ = 1, thus decreasing
After it was determined that the vaporizer body would contain a detonation, the
bolt stresses were investigated. If a detonation were to occur, the bolts holding the flanges
on the end of exposed to large tensile loads. The conservative the vaporizer would be
assumption that the bolts carry the entire tensile load was used. Bolt pre-loads typically
serve to reduce the stress experienced in a bolt, and in the actual build up would be
utilized.[15] However, to make the estimate conservative, bolt pre-load was not assumed
for the stress calculation. The bolt stress calculation is shown is Equation B.2
Equation (B.2)
The grade 8 steel bolts used to make the vaporizer have a proof strength of 120
ksi.[15] Bases on the stress calculated Equation B.2, the factor of safety in the bolts is 2.65
utilized. A bolt pre-load will reduce the stress on the bolts. Since this will increase the
factor of safety of the system, an appropriate pre-load was calculated. The relationship
Equation (B.3)
108
Where T is the bolt torque, d is the bolt diameter, K is the torque constant and Fi is
the bolt pre-load. The bolts used in the construction of the vaporizer are 0.3125 inch
bolts. The stress area of the bolt is 0.0581 square inches for a 0.3125-20 bolt. The stress
area for a bolt is different from the geometric cross-sectional area due to the threads.
Torque constant is mainly a function of the friction, and to a lesser extent the thread
geometry. The value quoted for non-plated black finish bolts without lubricant is
K=0.30[15]
The bolt preload is defined in Equation B.3, and is based on the maximum
Equation (B.4)
Equation (Equation B.4 yields a bolt preload of 2625 pounds. This bolt force can be
equated to a nut torque. During instillation, it is possible to measure the bolt torque
readily. Using Equation B.3, the required bolt torque is 20.5 ft-lbf. The final step is to
calculate the maximum allowable torque to ensure that 20.5 ft-lbf is within limits. The
maximum bolt torque is determined assuming that the bolt pre-load is 75% of the proof
load [15], and that grade 8 bolts with a proof strength of 120 ksi are used. Using the above
In order to hold the vaporizer together in the event of a detonation, the bolt preload must
be greater than the value calculated for the highest pressure case, yet smaller than the
maximum torque value calculated based on the proof load of the bolt. Since the
109
maximum pressure case yielded 20.5 ft-lbf and the maximum torque based on bolt
material constraints is 41 ft-lbf, the current bolt arrangement is sufficient to hold the
110