Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

AIAA SciTech Forum 10.2514/6.

2018-0400
8–12 January 2018, Kissimmee, Florida
2018 AIAA Aerospace Sciences Meeting

Parameter Impact on Heat Flux in a Rotating Detonation


Engine

Samuel J. Meyer 1 and Marc D. Polanka 2


Air Force Institute of Technology, Wright-Patterson Air Force Base, OH, 45433

Frederick R. Schauer 3
Air Force Research Labs, Wright-Patterson Air Force Base, OH, 45433

John L. Hoke 4
Downloaded by UNIVERSITY OF ADELAIDE -INTERNET on January 15, 2018 | http://arc.aiaa.org | DOI: 10.2514/6.2018-0400

Innovative Scientific Solutions, Inc., Dayton, OH, 45439

Rotating Detonation Engines (RDEs) currently require active cooling techniques to


dissipate the large amount of heat released by detonation waves. The heat flux generated has
been shown to be a function of several parameters. Characterizing the heat transfer response
to these parameters will aid in the ability to design a cooling system for the RDE that
minimizes the amount of active cooling required while ensuring the hardware will survive.
The heat flux was determined by placing high-frequency heat flux arrays at the surface of
RDE walls. This effort captured the effects on heat transfer from aerospike nozzles, channel
width, time from ignition, and equivalence ratio. Increasing the channel width to 22.9 mm
increased the heat flux compared to 16.3 mm channel widths. Aerospike nozzle installation
resulted in multiple waves occuring simultaneously during operation, although the bulk heat
flux was not altered significantly as a result. The local heat flux increased at lower axial
locations while it decreased at higher axial locations when the aerospike nozzle was installed
indicative of a shift in the detonation location. The bulk heat flux varied in time independently
of the time from ignition. Increasing the equivalence ratio between 0.6<φ<1.2 resulted in the
bulk heat flux increasing and increasing at a larger rate in the 22.9 mm channel compared to
the 16.3 mm channel.

Nomenclature
𝐴𝐴𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒 = Cross-Sectional Area at RDE Exit Plane
𝐴𝐴𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖 = Cross-Sectional Area at RDE Inlet Plane
AFRL = Air Force Research Laboratory
CFD = Computational Fluid Dynamics
C-J = Chapman-Jouguet
LHV = Lower Heating Value
PDE = Pulsed Detonation Engine
𝑞𝑞̇ " = Heat Flux
RDE = Rotating Detonatoin Engine
RTD = Resistance Temperature Detector
𝑇𝑇𝑠𝑠 = Surface Temperature

I. Introduction

D etonation propulsion is particularly interesting to the combustion community due to the resulting pressure gain
from reactants to products as well as the decreased entropy generation compared to traditional deflagration

1
Doctoral Student of Aeronautical Engineering.
2
Professor of Aeronautical Engineering and AIAA Associate Fellow.
3
Mechanical Engineer, Aerospace Systems Directorate, AFRL/RQTC, AIAA Associate Fellow.
4
Research Engineer, 7610 McEwen Rd, AIAA Associate Fellow.
1
American Institute of Aeronautics and Astronautics

Copyright © 2018 by the American Institute of Aeronautics and Astronautics, Inc.


All rights reserved.
combustors [1]. Pulsed Detonation Engines (PDEs) were the legacy detonation devices and were limited by their
requirement for a detonation to be reinitiated during every cycle and their maximum thrust being limited by operational
frequency [2]. Rotating Detonation Engines (RDEs) do not require a detonation to be initiated during every cycle, but
rather maintain a detonation wave traveling circumferentially around an annulus while fresh products are injected,
detonated, and then expanded and purged from the chamber further axially downstream. Additionally, the high
operational frequencies of RDEs allows the engine to provide greater thrust compared to PDEs.
However, the increased operational frequency, current lack of bypass flow, and large surface area to volume ratios
results in the surrounding structure absorbing more heat compared to stuctures in traditional deflagration combustors.
These elevated heat loads often limit the run times for laboratory combustors to seconds without active cooling. A
method of carrying heat away from the engine walls must be devised in order to operate RDEs for more practical
lengths of time. The first step requires understanding the local instantaneous heat flux within the RDE. Previous
investigations have revealed unsteady effects present within the RDEs of 2500-3700 Hz [3][4]. This heat flux has
also been shown to be dependant on several different parameters including mass flow rate, equivalence ratio, and
Downloaded by UNIVERSITY OF ADELAIDE -INTERNET on January 15, 2018 | http://arc.aiaa.org | DOI: 10.2514/6.2018-0400

channel width.

II. Background

RDEs operate by detonating a fuel/air mixture within an annulus. The injection geometry for fresh reactants must
be sized properly to ensure enough inflowing reactants refill the annulus and are adequately mixed prior to the
detonation wave completing its revolution. As seen in Figure 1, Schwer and Kailasanth [5] identified seven major
regimes within the RDE: the detonation wave (A), the oblique shock (B), and the shear layer separating the old and
fresh products (C), the detonation products (D), the layer between the detonation products and the inflowing reactants
(E), the high pressure region preventing inflowing reactants from entering the chamber (F) and the fresh reactants (G).

Figure 1. Typical RDE wave-following temperature solution with significant features identified [5]

The RDE has fuel and air flow into the channel in Region G. The inflowing reactants are consumed by a detonation
wave seen in Region A. The pressure of Region F is large enough that it prevents new reactants from flowing into the
chamber until the detonation products in Region D have expanded to a lower pressure than the inflowing reactants.
The products continue to expand and a barrier forms between the fresh, inflowing reactants and the products, seen in
Region E. An oblique shock is formed protruding from the detonation, seen in Region B which results in a shear layer
between the oldest and newest products, seen in Region C. The heat flux experienced in these seven regions will be
different. Region A has the largest heat flux because it exists at the detonation front where the heat is being released.
The heat flux decreased both circumferentially and axially with distance from Region A because the local temperature
of the flow decreased. Since the detonation is moving inside an RDE, the location of peak heat release also varied
around the annulus resulting in an spatially varying unsteady heat flux.
Only limited data are available in literature regarding attempts to measure heat flux from passing detonation waves.
Initial measurements by Bykovskii et al [6] were made at low frequency (1 kHz) and determined that the largest heat
flux values occurred at the axial height of the detonation wave. The refill region (Region G of Figure 1) and the region
2
American Institute of Aeronautics and Astronautics
axially downstream of the detonation wave experienced lower heat transfer. Bykovskii et al. postulated the cool refill
gas reduced the heat load below the detonation. The heat release downstream of the detonation was likely less due to
the gas temperature decreasing as a result of the expansion process. Interestingly, Bykovskii et al. altered the chamber
geometry to promote deflagration rather than detonation and determined the heat transfer to the wall was not
significantly different between the two modes [6].
Theuerkauf et al. [7] used a water-cooled RDE to make steady-state bulk heat flux measurements and a film-based
heat flux array with platinum wires functioning as an RTD deposited on both sides of the film to measure instantaneous
heat flux. These arrays had been previously used to measure high frequency heat flux on turbine blades [8] and
converted the measured signals to a heat flux using techniques described by Oldfield [9]. The heat flux arrays survived
for no more than 200 msec. Theuerkauf et al. found as much as 20% of the lower hearing value of the mixture was
transferred to the wall and that heat flux increased as mass flow rate increased and as equivalence ratio increased for
0.7<φ<1.2 [5]. Theuerkauf et al. warned that their bulk heat flux measurements could be manipulated based on the
coolant flow rate used (82 SLPM) since higher coolant flow rates are expected to remove more heat from the flow.
Downloaded by UNIVERSITY OF ADELAIDE -INTERNET on January 15, 2018 | http://arc.aiaa.org | DOI: 10.2514/6.2018-0400

In a separate study, Theuerkauf et al. were able to capture the first high-frequency (100 kHz) surface heat flux
waveforms on the wetted surface of an RDE [10]. They adhered an RTD array containing eight RTDs on each side of
a polyimide substrate to the outer body of an RDE and gathered data for high-frequency heat fluxes including
instantaneous heat flux at a fixed location on the wall. They also synchronized high speed chemiluminescence to
compare the waveform to the detonation location. They discovered the bulk heat flux was lower during initial startup,
but increased as the detonation wave became more stable. The captured waveform showed sharp peaks corresponding
to the detonation wave passing. Prior to these peaks, negative heat fluxes occurred that were attributed to the in
flowing reactants removing heat from the surface as postulated by Bykovskii [10]. However, Bykovskii et al. were
unable to measure the negative heat fluxes due to the lower frequency techniques they employed.
Theuerkauf et al. compared heat flux data gathered from experimentation to heat flux data collected from CFD
simulations performed by Dr. Paxson [3]. The CFD was a high-resolution algorithm integrating quasi-two-
dimensional, single species, reactive Euler equations with source terms while utilizing a course grid in the detonation
frame of reference. The CFD attempted to model the bulk heat fluxes at fixed axial locations above the fuel plate and
compared that data to bulk heat flux experimentally measured. The heat flux measurements were made by twelve
platinum elements connected by gold/copper leads placed on a glass-mica ceramic. The ceramic eliminated the
cratering that regularly occurred with the previously used polyimide sheets. The CFD predicted the axial location of
the maximum heat flux, however the CFD determined the magnitude of the heat flux to be about 40% higher than the
experimental value [3]. No explanation for the difference in magnitude was given.
Meyer et al. [4] used a twelve array gauge to measure the instantaneous heat flux and simultaneously measured
the temperature of the gas across the channel using TDLAS techniques. These two measurements, along with the wall
temperature from the array, allowed the high-frequency heat transfer coefficient to be determined for the first time.
Meyer et al. used the same RDE as Theuerkauf et al [3], but changed the outer body to alter the channel width from
7.6 mm to 16.2 mm. Meyer et al. determined that increasing the channel width resulted in fewer negative heat flux
measurements on the outer body and suggested a recirculation zone present in radial flow RDEs was responsible.
Additionally, the larger channel width resulted in a ~70% decrease in the heat flux to the outer wall. The heat transfer
coefficient had a larger bulk value and larger instantaneous value at ~80mm locations compared to ~20mm locations.
The equivalence ratio of the flow had a noticeable impact on the heat transfer coefficient at ~80mm as the heat transfer
coefficient grew by 350% from 0.7 to 1.1, while a variation in mass flow rate from 0.31 kg/sec to 0.62 kg/sec did not
have a significant impact on the heat transfer coefficient [4].
Randall et al. performed an experimental and numerical study of heat transfer in the RDE by placing thermocouples
at varying depths into sensor ports in the outer body and operated the RDE over a range of equivalence ratios and
mass flow rates for four second experiments using hydrogen-air [11]. Randall et al found that lower frequency
experimental runs experienced lower temperatures and their numerical study showed that the interior of the containing
walls (the portion interacting with the detonation) experienced higher temperatures than other portions of the wall
[11].
Roy et al. [12] also performed CFD simulations using 1-D and 2-D numerical models and compared their data
with previously published data from Randall et al [11]. Roy et al. predicted that outer body wall temperatures can rise
as much as 500 K after only 3.5 seconds of operation and reach a temperature of 1100 K within the detonation region.
Randall et al additionally stated that their model likely under-predicted the steady-state surface temperature due to the
impact of circumferial heat transfer and the questionable validity of the semi-infinite solid assumption for long
duration runs to steady-state. Roy et al. predicted the temperatures of the containing walls will be close to the melting
point of the material and thus require convective cooling of the combustor walls [12].

3
American Institute of Aeronautics and Astronautics
The primary goal of this experiment is to determine how the heat transfer is impacted by varying different
parameters. Specific parameters varied include the channel width, equivalence ratio, mass flow rate, and exit area
ratio. Previous heat transfer calculations by Meyer et al. [4] will be included for determining how the heat transfer is
impacted by different parameters.

III. Methodology
Primary testing for this experiment was performed at AFRLs DERF. The RDE seen in Figure 2 is a 139 mm inner
diameter annulus. The outer body of the RDE can be changed for altering the channel width. Fuel and air are delivered
separately into the chamber where the two streams mix within the detonation channel. A predetonator initiates the
detonation process. The detonation wave continually propagates around the annulus and the combustion products are
expanded as they proceed through and exit the annulus.
Tested fuels included hydrogen
Downloaded by UNIVERSITY OF ADELAIDE -INTERNET on January 15, 2018 | http://arc.aiaa.org | DOI: 10.2514/6.2018-0400

and ethylene. A range of airflows


from 0.31-0.62 kg/sec and
equivalence ratios from 0.6 to 1.2
were considered as well as channel
widths of 16.3 mm and 22.9 mm.
Additional tests were also run with
an aerospike nozzle added to the
RDE to reduce the exit area by 40%
in order to achieve an increased back
pressure for the environment.
A twelve sensor heat flux gauge
array similar to the arrays used by
Theuerkauf et al. [9] and Meyer et al.
[10] is shown in Figure 3. The array
consists of glass-mica ceramic
selected for its low conductivity, low
reactivity, easy accessibility, as well
as ease in manufacturing additional
future plates. The lower thermal
conductivity of the plate also allows
for longer potential run durations
prior to the semi-infinite solid
assumption no longer being valid.
Figure 2. Cross-section of AFRL 6-inch RDE Twelve platinum RTD’s were placed
onto the ceramic plates consisting of
thin platinum elements. The array in Figure 3 contains gold/copper alloy leads overlaying the platinum RTDs and was
similar to the one used for testing. Each RTD was connected individually to a Wheatstone quarter bridge was supplied
with a 30V potential difference. The output voltage from each bridge was amplified by a factor of 32-35 and passed
through a 330 kHz low-pass filter prior to data acquisition. Meyer et al. [10] have previously shown how this process
takes a measured voltage and converts it to surface temperature and heat flux.
The heat flux gauge was mounted to the outer body wall of the RDE. A gasket of 0.8 mm thickness silicone sealed
the plate against the steel wall of the outer body as well as ensuring no direct contact between the ceramic plate and
the outer body wall. The outer wall was 114 mm high with a diameter of 184 mm. The center body height was 105
mm with a 139 mm diameter. The air gap was 2.6 mm tall and the fuel plate consisted of 120 0.71 mm injection holes.
The RDE wall surface temperature was determined by the high-frequency temperature gauge. Over short time
scales, the high frequency temperature measurement was leveraged to provide a heat flux using techniques proposed
by Oldfield [13]. This technique involves solving for heat flux as a convolution integral of a response function and
the temperature measured on the Macor plate. Calibration on the plate was performed by immersing the plate in a
thermal bath and measuring the resistance change across the platinum leads as the temperature varied. The data was
plotted and a linear slope was fitted for each gauge with a typical 𝑅𝑅2 value of 0.9999 demonstrating a strong linear
relationship between the resistance of the gauge and the percent change in temperature it experiences. Overall,
theuncertainty in the temperture measurement is ± 1.2 K while the instantaneous heat flux has an uncertainty of ± 0.02

4
American Institute of Aeronautics and Astronautics
𝑀𝑀𝑀𝑀 ⁄𝑚𝑚2 . The temperature change of the macor plate during testing was within the temperature change covered by
the calibration

IV. Results
The targeted test case included multiple equivalence ratios between 0.6<φ<1.2 at a mass flow rate at 0.31 kg/sec
on a 22.9 mm channel. Previous
measurements at the same mass
flow rate were made on 7.6 mm
and 16.3 mm channels [3][4].
Additionally, an aerospike nozzle
was installed with a 0.6 area ratio
Downloaded by UNIVERSITY OF ADELAIDE -INTERNET on January 15, 2018 | http://arc.aiaa.org | DOI: 10.2514/6.2018-0400

(𝐴𝐴𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒 ⁄𝐴𝐴𝑖𝑖𝑖𝑖𝑖𝑖𝑒𝑒𝑡𝑡 ) to examine its


effects at different equivalence
ratios. Two separate plates were
used in this testing. The first plate
was used for most test cases
including all of the aerospike
nozzle cases. Two gauges locted
at 24.5 mm and 39.5 mm axially
were primarily used on this plate.
The 24.5 mm gauge was expected
to be within the detonation region
for most test cases, while the 39.5
mm gauge was expected to be
Figure 3. Twelve-sensor heat flux gage array on glass mica ceramic above the detonation. The second
substrate plate used contained gauges at 9.7
mm, 24.5 mm, 39.5 mm, 69.1
mm, and 76.4 mm. For identical test cases and equivalence ratios, the two plates measured bulk heat fluxes within 5%
of each other which is well within the variability for heat flux shown in Figure 4. All test cases achieved detonation.
Figure 4 shows the variation in heat flux experienced by plate two for the five axial locations. Notably, the heat
flux on the 22.9 mm channel shows a large cycle to cycle deviation between 30-55% compared to previously published
results for the 16.3 mm channel that were closer to 2-15% [4]. This suggests that that heat flux between individual
revolutions varied significantly around the mean
bulk value. This was possibly caused by the
detonation moving radially within the RDE
throughout a revolution. The axial heat flux profile
seen in Figure 4 was the typical shape of the profile
and typical of the magnitude of variation
experienced in other test cases.
Figure 5 shows a common shape of the
waveform seen for hydrogen-air detonations at 0.31
kg/sec for an equivalence ratio of 0.8. The sharp
spikes in the heat flux represent the single
detonation wave passing by the gauge location and
was followed by a rapid decline in the heat flux.
This waveform shape noticeably differs from
previously published heat flux waveforms [9] [10].
Although the heat flux peaks were not stationary in
previous experiments, in this experiment, they
varied more noticeably. This variation in
magnitudes between successive revolutions suggests
that the increased channel width allows the
Figure 4: Bulk Heat Flux vs. Axial Distance for
detonation to have an additional freedom to move
Hydrogen-Air, 𝒎𝒎̇=0.31 kg/sec φ=0.66
radially within the channel compared to the smaller

5
American Institute of Aeronautics and Astronautics
channels previously investigated [9][10]. This corresponded to an instantaneous heat flux with greater unsteadiness
compared to smaller channels.
Downloaded by UNIVERSITY OF ADELAIDE -INTERNET on January 15, 2018 | http://arc.aiaa.org | DOI: 10.2514/6.2018-0400

Figure 5: Heat Flux vs. Time for Hydrogen-Air, 𝒎𝒎̇=0.31 kg/sec φ=0.80, Measured at 24.5 mm Above the Fuel
Plate

Figure 6 plots the bulk heat flux within 22.9 mm channel compared to the 16.3 mm channel across a range of
equivalence ratios. For the 22.9 mm channel, only the gauges located at 24.5 mm and 39.5 mm were used for these
calculations as these were the only two gauges acquired data during every test case. For the 16.3 mm channel, the heat
flux was averaged over multiple gauges from 15 mm - 80 mm axial location. From these previous measurements, the
gauges at 24.5 mm and 39.5 mm were expected to be within 2% of the bulk heat flux across the 15 mm – 80 mm axial
distance [4]. Previously, the heat flux was observed to decrease by between 73% - 82% by increasing the channel
width from 7.6 mm to 16.3 mm. Here, increasing the channel width to 22.9 mm did not result in further decreases in
heat flux, but instead caused in an increase at the outer wall. The difference in heat flux between the two cases was
minimal at the leaner equivalence ratios and became more pronounced as the equivalence ratio trended towards rich
conditions. Increasing the channel width while maintaining the massflow rate results in the mass flux through the
engine decreasing resulting in a smaller Reynolds number as well as a smaller Nusselt number. The smaller Nusselt
number results in a smaller convective heat transfer coefficient. In order to attain comparable or larger heat fluxes
with a smaller convective heat transfer coefficient, the adiabatic wall temperature must be larger. Possible
explanations for this trend is that the detonation, while moving radially with every revolution, was actually closer on
average to the outer wall for the 22.9 mm channel than the 16.3 mm channel, or perhaps a high temperature
recirculation region was present at the outer body preventing cool refresh reactants from interacting with the wall.
The next test points consisted of installing an aerospike nozzle on the RDE with a 0.6 Area Ratio (𝐴𝐴𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒 ⁄𝐴𝐴𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖 ).
The aerospike nozzle had several effects on the flow within the RDE. First, the aerospike nozzle lessened the area for
the products to exit which caused the pressure inside the channel to increase. This increase in pressure does not allow
the incoming reactants to penetrate the same axial depths. This in turn caused the detonation height to be lowered
which was expected to increase the heat flux at axial locations lower in the channel. Additionally, the aerospike nozzle
forced the flow to exit the RDE near the outer body of the RDE while directing flow away from the inner body of the
RDE. It was expected that this caused heat flux measurements at higher axial locations to be increased compared to
similar measurements without an aerospike nozzle installed. The plate used for these testing had no working gauges
at a high enough axial location to quantify this effect. Figure 7 shows the bulk heat flux measured at axial heights with
24.5 mm and 39.5 mm. A curvefit of the points between 0.7<φ<1.0 shows the aerospike nozzle increased the heat flux
at the 24.5 location by ~14% while decreasing the heat flux at the 39.5 mm location by ~6% between 0.7<φ<1.0. This
is consistent with a change of axial location of the detonation to a lower postion and thus the heat release. Overall,
this result suggests that the aerospike may manipulate the location of the heat release and thus the peak heat flux
axially, but may not change the total heat flux across the entire outer body. Further testing with additional working
axial gauges needs to be accomplished to make that determination.
An additional effect that can occur when the pressure inside an RDE increases is multiple detonation waves. As
pressure inside the RDE increases, the detonation cell width and length decreases [14] causing the propagation velocity
of the detonation wave to decrease. In order to consume all of the inflowing reactants, additional detonation waves
may develop inside of the RDE. When the aerospike nozzle was installed, multiwave operation occurred for test points
with an equivalence ratio of 0.79 and above. The presence of the multiwave accounts for part of the changes seen in
Fiugre 7. Having two waves for 0.79<φ<1.0 created a change in the trends between the two gauges shown. A slight
increase in heat flux occurred for the lower gauge and a corresponding decrease happened for the downstream gauge.
6
American Institute of Aeronautics and Astronautics
Downloaded by UNIVERSITY OF ADELAIDE -INTERNET on January 15, 2018 | http://arc.aiaa.org | DOI: 10.2514/6.2018-0400

Figure 6: Comparison of Heat Flux at the Outer Body of an RDE for Different Channel Widths

Figure 7: Effect of Aerospike Nozzle on Heat Flux for Different Equivalence Ratios

7
American Institute of Aeronautics and Astronautics
This again is consistent with the heat release occurring at a lower axial postion. Multiwave operation was verified
by observing the fourier transform of the data with the primary frequency occuring between 3.9-4.1 kHz, which is a
higher frequency than the Chapman-Jouguet (C-J) velocity (C-J frequency between 3.65-3.95 kHz depending on φ).
Since the detonation wave cannot achieve a velocity higher than the C-J velocity, the only possibility is the presence
of multiple detonation waves within the engine. Multiple waves appear to develop within the first five milliseconds
of operation for the three points that achieved multiwave operation.
Figures 8 and 9 shows typical heat flux waveforms seen for multiwave operation. Figure 8 shows the typical shape
of a multiwave waveform where a greater unsteadiness was experienced than the waveform seen in Figure 4 with
more peaks and troughs. This suggests that multiwave operation will likely lead to enhanced thermal fatigue compared
to single wave operation. The heat flux peaks, however, are not as high in multiwave operation compared to single
wave operation for the similar equivalence ratios tested. Although multiwave operation consists of multiple detonation
waves which are presumed to increase the heat flux, the heat flux peaks being notably lessened causes the bulk heat
flux to be comparable for the time scales considered. Figure 9 shows a multiwave waveform where one of the two
Downloaded by UNIVERSITY OF ADELAIDE -INTERNET on January 15, 2018 | http://arc.aiaa.org | DOI: 10.2514/6.2018-0400

detonation waves has a noticeably higher heat flux than the other. Potential reasons may be one of the detonations is
stronger than the other or because each detonation may be moving inward and outward radially with every revolution.
Additionally, one detonation spike being routinely larger in multiwave operation will have an impact on the thermal
fatigue.

Figure 8: Waveform of Multiwave Operation Taken for Hydrogen-Air, 𝒎𝒎̇=0.31 kg/sec, φ=0.96, Measured at
24.5 mm Above the Fuel Plate

Figure 9: Waveform of Multiwave Operation with One Wave Appearing Stronger Than the Other for
Hydrogen-Air, 𝒎𝒎̇=0.31 kg/sec, φ=0.96, Measured at 24.5 mm Above the Fuel Plate

Lastly, the secondary goal of this research was to investigate how the instantaneous heat flux changes for longer
duration runs. As time progresses the wall temperature achieves a higher temperature which is expected to lessen the
heat flux. Previous attempts to capture instantaneous heat flux have focused on run durations under 200 milliseconds.
Test points accomplished for this test attempted run durations of over 250 milliseconds. Figure 10 shows the
instantaneous heat flux from 74-76 milliseconds and 274-276 milliseconds. The waveforms show similar maximum
peak heat fluxes and display a large variability in peak heat flux between detonations, but there does not appear to be
a noticeable difference between the two waveforms. Figure 11 shows the bulk heat flux during ten millisecond

8
American Institute of Aeronautics and Astronautics
increments as a function of time. The first 30 milliseconds were omitted to only study results after the startup
transients. The bulk values appear to diminish by 10% after the first 90 milliseconds, however bulk heat flux swings
as great as 15% between 10 millisecond intervals were seen in several cases not shown here. This led to the conclusion
that the bulk heat flux was not impacted by time from detonation initiation for the 30-300 millisecond timescales
considered for all test cases observed. The result in Figure 11 is consistent with results from other equivalence ratios.
Downloaded by UNIVERSITY OF ADELAIDE -INTERNET on January 15, 2018 | http://arc.aiaa.org | DOI: 10.2514/6.2018-0400

Figure 10: Singlewave Heat Flux Waveform From 74-76 msec and 274-276 msec for Hydrogen-Air,
𝒎𝒎̇=0.31 kg/sec, φ=0.80, Measured at 24.5 mm Above the Fuel Plate

Figure 11: Bulk Heat Flux vs. Time for Hydrogen-Air, 𝒎𝒎̇=0.31 kg/sec, φ=0.80

V. Conclusions

The instantaneous and bulk heat fluxes impact from multiple parameters were considered in this investigation. The
variables altered included the channel width, presence of an aerospike nozzle, and equivalence ratio. Interestingly,
increasing the channel width from a previously tested 16.3 mm to 22.9 mm resulted in a 12% increase, on average,
of heat flux to the outer wall for the 22.9 mm channel. The difference in heat flux for the two channel widths appears
to be negligible at leaner equivalence ratios, but increases as the mixture becomes richer. The cause of this increase
was conjectured to be a change in the position of the detonation to a location closer to the outer wall in the larger
channel. Adding an aerospike nozzle to the engine did not impact the total bulk heat flux through the outer wall, but
it did impact the axial location of the heat flux. The bulk heat flux appears to increase by about 14% at 24.5 mm axial
locations while decreasing by about 6% at 39.5 mm axial locations. Multiwave operation occurred in three test cases
9
American Institute of Aeronautics and Astronautics
marking the first time the authors are aware of that an instantaneous heat flux waveform was measured for multiwave
operation. The waveform appears to be more unsteady which may lead to thermal fatigue occuring quicker during
multiwave operation. During multiwave operation, the peak instantaneous heat flux decreased compared to singlewave
operation. However, multiwave operation showed very little difference in total bulk heat flux compared to single wave
operation. At times during multiwave operation the heat flux spike for each individual wave can have noticeably
different magnitudes suggesting that during multiwave operation one of the detonations may be stronger than the
other. For all test cases considered, the bulk heat flux did not vary in a consistent manner with time from ignition for
the first 300 milliseconds but the bulk heat flux value for isolated 10 millisecond increments could vary by as much
as 15% of the time averged bulk value.

Acknowledgements
The authors wish to thank Dr. Richard Anthony for assistance with the heat flux plates as well as Mr. Curtis Rice
Downloaded by UNIVERSITY OF ADELAIDE -INTERNET on January 15, 2018 | http://arc.aiaa.org | DOI: 10.2514/6.2018-0400

for his assistance operating equipment within the laboratory as well as ensuring a safe setup and experiment took
place.

References
1. Heiser, W., and Pratt, D., “Thermodynamic Cycle Analysis of Pulse Detonation Engines” Journal of
Propulsion and Power, Vol. 18, No. 1, January-February 2002, pg 68-76.
2. Schauer, F., and Bradley, R., “Detonation Initiation Studies and Performance Results for Pulsed Detonation
Engine Applications,” 39th AIAA Aerospace Sciences Meeting, AIAA 2001-1129, Reno, NV, 2001.
3. Theuerkauf, S., Schauer, F., Anthony, R., Paxson, D., Stevens, C., and Hoke, J., “Comparison of Simultated
and Measured Instantaneous Heat Flux in a Rotating Detonation Engine”, 54th AIAA Aerospace Sciences
Meeting. AIAA 2016-1200, San Diego, CA, 2016
4. Meyer, S. J., Polanka, M. D., Schauer, F. R., Anthony, R. J., Stevens, C., Hoke, J. L., and Rein, K.,
“Experimental Characterization of Heat Transfer Coefficients in a Rotating Detonation Engine,” 55th AIAA
Aerospace Sciences Meeting, AIAA 2017-1285, Grapevine, TX, 2017.
5. Schwer, D.A. and Kailasanth, K, “Numerical Investigation of Rotating Detonation Engines,” 46th
AIAA/ASME/SAE/ASEE Joint Propulsion Conference and Exhibit, AIAA 2010-3880, Nashville, TN, 2010.
6. Bykovskii, F.A. and Vedernikov, E. F., "Heat Fluxes to Combustor Walls during Continuous Spin Detonation
of Fuel-Air Mixtures", Combustion, Explosion, and Shock Waves, 44: 70-77 (2009).
7. Theuerkauf, S., Schauer, F., Anthony, R., and Hoke, J., "Average and Instantaneous Heat Release to the
Walls of an RDE", 52nd AIAA Aerospace Sciences Meeting. AIAA 2014-1503, National Harbor, MD, 2014
8. Anthony, R. J., Oldfield, M. L. G., Jones, T. V., LaGraff, J. E., “Development of High-Density Arrays of
Thin Film Heat Transfer Gages,” AJTE99-6159, Proceedings of the 5th ASME/JSME Thermal Engineering
Joint Conference, San Diego, CA, 1999.
9. Theuerkauf, S., Schauer, F., Anthony, R., and Hoke, J., "Average and Instantaneous Heat Release to the
Walls of an RDE", 52nd AIAA Aerospace Sciences Meeting. AIAA 2014-1503, National Harbor, MD, 2014
10. Theuerkauf, S., Schauer, F., Anthony, R., and Hoke, J., “Experimental Characterization of High-Frequency
Heat Flux in a Rotating Detonation Engine”, 53rd AIAA Aerospace Sciences Meeting. AIAA 2015-1603,
Kissimmee, FL, 2015
11. Randall S., St. George, A., Driscoll, R., Anand, V., Gutmark, E,J., “Numerical and Experimental Study of
Heat Transfer in a Rotating Detonation Engine”, 53rd AIAA Aerospace Sciences Meeting, AIAA 2015-0880,
Kissimmee, FL, 2015.
12. Roy, A., Strakey, P., Sidwell, T., and Ferguson, D., “Unsteady Heat Transfer Analysis to Predict Combustor
Wall Temperature in Rotating Detonation Engine”, 51st AIAA/SAE/ASEE Joint Propulsion Conference.
AIAA 2015-4191, Orlando, FL, 2015
13. Oldfield, M. L. G., "Impulse Response Processing of Transient Heat Transfer Gage Signals," Journal of
Turbomachinery, 130 (2008).
14. Stevens, C. A., Hoke, J. L., and Schauer, F. R., “Propane/Air Cell Size Correlation to Temperature and
Pressure,” 54th AIAA Aerospace Sciences Meeting, AIAA 2016-1400, San Diego, CA, 2016.

10
American Institute of Aeronautics and Astronautics

You might also like