You are on page 1of 10

Experimental Thermal and Fluid Science 102 (2019) 261–270

Contents lists available at ScienceDirect

Experimental Thermal and Fluid Science


journal homepage: www.elsevier.com/locate/etfs

Diagnostic method based on spontaneous emission to evaluate the T


detonation cycle
Nathan C. Lopesa, Carlos C.B. Katataa, Carla S.T. Marquesa,b,

a
Aerothermodynamics and Hypersonics Division, Institute for Advanced Studies – Department of Aerospace Science and Technology, Trevo Cel Av. José A. A. do Amarante,
n° 1, 12228-001 São José dos Campos, SP, Brazil
b
Technological Institute of Aeronautics – Department of Aerospace Science and Technology, Praça Mal Eduardo Gomes, 50, 12228-900 São José dos Campos, SP, Brazil

ARTICLE INFO ABSTRACT

Keywords: A study on a hydrogen-fuelled single-cycle pulse detonation device, through pressure and spontaneous emission
Spontaneous emission measurements, was carried out. Detonation initiation was effectively achieved by applying a 65 J microsecond
Detonation discharge associated with the Shchelkin spiral with very small errors in the diagnostic method. OH* and H2O*
Optical diagnostics temporal emission measurements were applied at the nozzle exit to establish correlations to specific impulses,
OH*
determined from pressure curves. OH* emission images were also captured to verify detonation propagation at
PDE
the exhaust. Both the OH* and H2O* emitters correlated well with the specific impulses, and the integrated areas
under emission peak maxima resulted in improved correlations. It was found that the OH* amount per kilogram
of fuel grows exponentially with the specific impulse based on fuel, while the H2O* emission area per kilogram
of mixture decays exponentially with the specific impulse, based on the mixture. The OH* images demonstrate
stable detonation propagations and validate the established correlations. Furthermore, a new optical method
was developed for detonation cycle diagnostics.

1. Introduction It is intended that PDEs will replace current engines for truck power,
electricity generation, navy ships, aircraft and rockets. They could be
The application of detonation for propulsion purposes in a systematic applied as isolated or integrated propulsion systems, and those that can be
form has been investigated since the mid-1950s due to the higher ther- driven to high speeds and have access to space are of particularly sig-
modynamic efficiency promoted by detonation cycles, with high pres- nificant interest [14,15]. Because of the high thrust and low weight of
sures and temperatures generated, than that of deflagrations [1,2]. The PDEs, a 25% reduction in fuel consumption could be obtained. In addi-
first detonation concept appropriate to propulsion was the oblique deto- tion, low cost, reduced complexity and an extensive flight regime make
nation wave engine (ODWE) for hypersonic flight [3]. However, the first these engines the aviation propulsion systems of the next generation [16].
propulsion system based on detonation for producing thrust was the pulse Various challenges still exist in terms of the full development of these
detonation engine (PDE) filled with H2-air mixtures, proposed and built at engines, mainly concerning achieving a controlled, reliable and repeatable
the University of Michigan by Nicholls et al. in 1957 [4,5]. detonation at a minimum length. The detonation initiation is the key issue
The renewed interest in PDEs during the 1980s resulted in nu- for an operational PDE, as this determines the engine chamber size and
merous experimental and computational studies, and advances cycle time and has a direct impact on working frequency. Furthermore, for
achieved in this topic made practical propulsion devices using deto- improved performance, they should operate at a high frequency [17,18].
nation a reality; therefore, these engines have attracted research at- Several PDE types have been considered for analysis, including the
tention across the entire world [4,6,7]. In 2008, the first flight of an very simple single-tube/single-cycle to multi-tubes/multi-cycles, val-
aircraft powered by a PDE, with four tubes operating at 20 Hz, was veless or with mechanical valves, with dimensions from a micro-engine
performed successfully by the US Air Force Research Laboratory [8]. (as igniter) up to very large systems [4,17,19–26].
Engine manufacturers such as GE [9,10], Pratt Whitney [11], Rolls Experimental initiation methods are aspects of significant research
Royce [12] and Volvo Aero [13] have pursued PDE technology and the that aim to improve PDE performance by decreasing cycle time. Direct
first two companies intend to produce them commercially in mid-2020. initiation of detonation is the most effective ignition method because it


Corresponding author at: Research Centre for Gas Innovation at the University of São Paulo – Department of Mechanical Engineering at the Polytechnic School,
Av. Prof. Mello Moraes 2231, 05508-030 São Paulo, SP, Brazil.
E-mail address: carla271170@gmail.com (C.S.T. Marques).

https://doi.org/10.1016/j.expthermflusci.2018.12.003
Received 25 April 2017; Received in revised form 1 September 2018; Accepted 2 December 2018
Available online 04 December 2018
0894-1777/ © 2018 Elsevier Inc. All rights reserved.
N.C. Lopes et al. Experimental Thermal and Fluid Science 102 (2019) 261–270

Fig. 1. Experimental setup for the acquisition of pressure and light emission measurements in detonation wave propagation.

eliminates the deflagration-to-detonation transition (DDT) and, there- Natural light emissions are simple optical measurements that are ap-
fore, shortens the cycle time. This can be realized through laser ignition propriate for following combustion chemistry and provide various other
[27], nanosecond plasma ignition [28–32], pre-detonator ignition features. In general, in PDEs, high-speed emission images have been ac-
[33–35] and the dual PDE crossover system [36]. Although optical and quired for the visualization of propagation waves (DDT and detonation)
transient plasma ignitions are promising technologies, they are hard to [17,19,28,30,43,47,48], reaction zones [17], heat release rate [47] and
handle, and due to their complex systems, they are still problematic to combustion efficiency [28,30]. Temperature distribution [50,51] and igni-
implement practically. Moreover, pre-detonator ignition, which is tion delay times [28,30,54] have also been determined by emission mea-
usually applied to rocket engines, requires a high pressure with O2 surements in these engines, and half of them have used filtered emissions to
mixtures on board, resulting in a more hazardous system. select a single emitter properly. Emitter radicals are exceptional reaction
Other ignition methods for PDE detonation initiation are multiple markers due to their reduced lifetimes, while thermal emitters are suitable
sparks [18,34,37,38] with or without obstacles, hot jet ignition for determining temperature. Both the emitter types could be employed as
[33,39,40] and low-energy ignition (as with automotive ignition) asso- probes of the heat release rate and combustion efficiency.
ciated with Shchelkin spiral-like-obstacles [10,12,15,17,19,20,25,41,42]. In this work, the expansion of H2-air detonations in the single-cycle
The last ignition method has been the most employed and was selected pulse detonation device at the Institute for Advanced Studies (IEAv/
for this work. Obstacles enhance the turbulence, which increases the DCTA) is characterized and evaluated through pressure measurements
flame-burning rate due to the augmented flame surface area and gra- as well as spontaneous OH* and H2O* emissions. Correlations between
dients in the mixture. However, it has been reported that an optimized emission time profiles and specific impulses are proposed, and emission
obstacle configuration is necessary for reducing DDT length and time images from the exhaust are captured to identify the narrow zone re-
without performance loss [17,43] because impulse measurements are action resulting from detonations.
significantly lower in tubes with obstacles than those without [42].
In PDEs, the mixture is immediately compressed by a shock wave, 2. Experimental setup
burning at a high pressure and constant volume; hence, an intermittent
detonation wave is produced and propagated to generate thrust. The The single-cycle pulse detonation device produced at the IEAv was
PDE thermodynamic cycle (Fickett–Jacobs cycle) is similar to the designed for experimental simulation of ideal combustion conditions of
Humphrey cycle but slightly more efficient. In spite of the similarities a PDE cycle with the main purpose of the application and development
between detonation and constant volume combustions, the highest PDE of optical diagnostics. Each experimental test ideally can represent the
efficiencies are directly related to the lower entropy growth of first detonation cycle of a PDE. The combustion processes involved in
Chapman–Jouguet detonations [44]. PDEs are very similar to those in scramjet engines, and characteristic
Optical diagnostic techniques have been widely applied to engines acquisition and detection are required to monitor high-speed combus-
for characterization and evaluation because they provide high resolu- tion, as occurs in these engines.
tion in space and time. In particular, they have been employed for PDEs The experimental device consists of an ignition system, 304 stainless
to establish and visualize shock waves, DDT, reaction zones, tempera- steel detonation tube, divergent nozzle and test chamber. A diaphragm
ture and concentrations [17,45–53]. Schlieren, planar laser-induced separates the explosive mixture in the detonation tube from the other
fluorescence, spontaneous emission and tunable diode laser absorption sections of the system under vacuum. Silica-fused windows in both the
are the most common optical diagnostics used for detonation analysis in detonation tube and test chamber allow optical measurements from
these engines. Optical sensors and methodologies are fundamental to ultra-violet to near-infrared. Fig. 1 shows the experimental setup for the
practical PDE development, which requires control for safe operation. acquisition of pressure and light emission measurements.

262
N.C. Lopes et al. Experimental Thermal and Fluid Science 102 (2019) 261–270

The explosive mixtures of H2-air were prepared using aluminized


Mylar film of 50 µm. Stoichiometric (Φ = 1.0) and fuel-rich (Φ = 1.2)
mixtures1 were placed in the detonation tube at 760 mmHg2 after 2 h
under vacuum, with the detonation tube at ca. 5 Pa and test chamber at
ca. 10 Pa. The gaseous mixtures were perfectly stirred for 30 min prior
to ignition. The mixtures’ partial pressures were controlled by digital
manometers, and the vacuum pressures were controlled by a digital
vacuometer. The detonation tube was filled with 225 ± 2 mmHg of H2
and 535 ± 5 mmHg of air for the stoichiometric mixtures3 and with
255 ± 2 mmHg of H2 and 505 ± 5 mmHg of air for the fuel-rich
mixtures4.
To achieve detonation initiation, a microsecond high-voltage pulse
of ca. 25 kV and 65 J was applied to a tungsten spark plug (NGK,
BUHW-2) with and without a Shchelkin spiral. The high-voltage pulse
duration was approximately 75 µs from a capacitive–inductive dis-
charge switched by a silicon-controlled rectifier (SCR). The Shchelkin
spiral was attached to the spark plug flange and installed in the first/
second sections of the detonation tube. Three spirals, with a blockage
( )
2
ratio (BR) of 0.44, BR = 1 [41], inner diameter of
dShchelkin
ddetonation tube
27 mm, and external diameter of 35 mm, and lengths of 330, 380, and
580 mm, were tested. The spirals were made of 102 steel wires of 4 mm
in diameter, and a pitch of 30 mm for those with a 330 mm length, and
36 mm for the others.
Subsequently, after ignition of the H2-air mixture, a combustion wave
propagates throughout the detonation tube, bursts the Mylar diaphragm,
Fig. 2. Tungsten spark plugs. Left: BUHW-2 from NGK (88% W, 9% Ni and 3%
and expands through the nozzle to the test chamber (Fig. 1). A digital
Cu alloy) with a gap of 1.8 mm, and right: home-made tungsten (70% W and
delay generator (SRS, DG645) was used to trigger the high-voltage pulse
30% Cu alloy) spark plug with a gap of 3 mm, and both electrodes lengthened.
and oscilloscope with eight channels (YOKOGAWA, DL 7450) simulta-
neously. The combustion wave propagation was followed by means of
piezoelectric pressure transducers (KISTLER, 603B), and the velocities scale by a home-made routine, to enable comparison. The images were
were determined in four tube regions with the same cross-section. later processed using GIMP2 software with a G’MIC plugin, for
Pressure transducers were located at 120 mm (P1), 190 mm (P2), smoothing and reducing data noise.
1295 mm (P3), 1365 mm (P4) and 1502 mm (P5) in the detonation tube
from the ignition start point. The sixth pressure transducer (P6) was
3. Results and discussion
placed at 1550 mm from the ignition start point, at the beginning of the
expansion region, 27 mm before the divergent nozzle. Light emission
The results of the microsecond spark ignition without the Shchelkin
signals were obtained by applying a fused silica plano-convex lens
spiral show a degraded detonation, with approximately 1000 m/s. A deto-
(ϕ = 76.2 mm and f = 250 mm), split by a ¼ m monochromator (ORIEL,
nation initiation was observed to decelerate at the end of the 1520 mm
Cornestone 260) and detected by a photomultiplier (HAMAMATSU,
length detonation tube. Initial velocities of 1038 ± 26 m/s were de-
R928) immediately at the nozzle exit. Time-resolved light emissions from
termined by applying an NGK tungsten spark plug, and much lower velo-
OH* (λ = 306.4 ± 1.05 nm) and H2O* (λ = 651.7 ± 2.1 nm) were
cities were observed at the exit of the detonation tube (533 ± 17 m/s) for
acquired throughout the detonation expansion.
stochiometric H2-air mixtures. For improved detonation initiation, a home-
Emission profiles were employed to establish the delay time for a
made tungsten spark plug, as shown in Fig. 2, was applied.
single image acquisition. An intensified charge-coupled device (ICCD)
However, in spite of improved initial velocities (1180 ± 20 m/s)
camera (PCO, Dicam-Pro) with a UV lens from NIKON (105 mm, f/4.5)
and lower deceleration, this was not sufficient to achieve sustained
was used to capture emission images, placed instead of the mono-
detonation for the entire tube length. Although higher ignition energies
chromator in Fig. 1, and focused on the nozzle exit. The focus was adjusted
could be employed, increased electromagnetic noise would then be
by taking visible images with a neutral density filter from the nozzle and
produced, and the ignition system could fail due to the limited electrical
window flange of the test chamber, illuminated by a mini Maglite flash-
isolation, as well as the spark plug being damaged by oxidation and
light. The entirety of the nozzle diameter, the window bottom and hor-
isolation rupture. Therefore, the standard tungsten spark plug (BUHW-
izontal window diameter were in the field of view for all images captured.
2) was applied with Shchelkin spirals to establish the detonation waves.
The field of view was limited by the relative aperture needed for images
Studies on DDT in stoichiometric H2-air mixtures have verified that
from high-speed flows. Furthermore, an interference filter was applied to
velocities above 1300 m/s at a 350 mm length successfully promote the
select the emission bands appropriately from OH*, and was centred at
transition to detonation. Obstacles such as the Shchelkin spiral could
307 ± 6 nm (FWHM = 12 nm) with a maximum transmittance of 12%.
significantly reduce the DDT length and time, which depends on the
H2O* images were also obtained, but they were not suitable for our pur-
blockage ratio (BR) and chemical energy released from the combustible
poses due to their large time and intensity fluctuations.
mixture [41]. Despite extreme difficulty in their optimization, with
The OH* emission images acquired were pre-normalized to the same
minimal BR and length associated with device size being desirable for
pixel value of the image with the highest maximum, and the same RGB
superior performance, this ignition method easily leads to detonation
conditions. Usually, overdriven and overpressure detonations are es-
1 (P /P
= HS2. R .air ,
)
where S.R, is the stoichiometric ratio for tablished [41,43].
2H2 + O2 + 3.76N2 = 2H2O + 3.76N2 reaction. The blockage ratio selected (BR = 0.44) is an optimal configuration;
2
In S.I. units 101.325 kPa however, it is determined for hydrocarbon fuels [42,55]. Three spiral
3
PH2 30.0 ± 0.3 kPa and Pair 71.3 ± 0.7 kPa lengths were tested with the aim of improving the diagnostic method
4
PH2 34.0 ± 0.3 kPa and Pair 67.3 ± 0.7 kPa because there are upper changes to specific impulses for different

263
N.C. Lopes et al. Experimental Thermal and Fluid Science 102 (2019) 261–270

14 Time at P1 (120 mm) - τDDT


Pressure at P1 (x 3) Start time at P5 (1502 mm)
12
Pressure at P2 (x 3) End time at P5 (1502 mm)
Pressure at P3 Start time at P6 (1550 mm)
Pressure at P4 Time of pressure peak at P6 (1550 mm) Φ = 1.0
10 5.2
5.1
8
5.0
Voltage (V)

4.9
6
4.8
4 4.7

Time (ms)
4.6
2
2.45

0 2.40
2.35
-2 2.30
3.2 3.4 3.6 3.8 4.0 4.2 4.4 4.6 4.8 5.0 5.2 2.25
Time (ms) 2.20
12 2.15
Pressure at P5 2.10
10
Pressure at P6 (x 3) 1 2 3 4 5 6 7 8 9 10 11
Peak 3 OH* emission Experimental Test
H O* emisssion
2
8
OH* start time
Time at peak 3 - τcycle H2O* start time
6 Peak 4 Peak Time at peak 4 Time at peak
Total time of peak 3 Total time of peak
Voltage (V)

Total time of peak 4 Total emission time


4
Total emission time
9.5
9.0
2 8.5
8.0
0 7.5
7.0
6.5
-2 6.0
5.5
5.0
Time (ms)

-4 4.5
4.4 4.6 4.8 5.0 5.2 5.4 5.6 5.8 6.0 6.2 6.4 4.0

Time (ms) 1.50


1.25
Fig. 3. Pressure curves and emission profiles for stoichiometric H2-air detona-
1.00
tion ignited by a microsecond discharge associated with a Shchelkin of 580 mm
0.75
length.
0.50
0.25 Φ = 1.0
spirals depending on the fuel percentage added. 0.00
The results showed that the Shchelkin with a 330 mm length did not 1 2 3 4 5 6 7 8 9 10 11
result in a detonation wave, whichever spark plug was used. Applying a Experimental test
380 mm length, velocities of approximately 1000 m/s were found in the
Fig. 4. Times from pressure curves and times from OH* and H2O* time profiles
entire detonation tube with pressures of about 1165 kPa. Therefore,
for each experimental test of the stoichiometric H2-air detonation.
only those with a 580 mm length resulted in sustained detonations.
Fig. 3 displays the typical results obtained for H2-air detonations
t0+ t
initiated by 65 J microsecond ignition associated with a Shchelkin of I= (P (t ) Pa) dt
t0 (1)
580 mm length.
Using pressure histories and emission profiles such as those in Fig. 3, The start time of the P6 transducer indicated the beginning of the
the time features of detonation waves, their velocities, and pressures at detonation expansion. The first light emissions from OH* and H2O* are
the end of the detonation tube were established as presented from now both usually ca. 0.1 ms after the start times at P6 (Figs. 4 and 5), and
on. exhibit pressure decays subsequent to zero pressure, as in Fig. 3. Hence,
Figs. 4 and 5 show the time features from pressure curves and from the light emissions follow the detonation expansion, not only on the
light emission profiles of the H2-air detonation waves. thrust surface where the emissions are maximized but from a certain
The time at the P1 pressure transducer (Figs. 4 and 5) is the ap- location in the first third part (1/3) of the expansion region until the
proximate detonation initiation time or tDDT because a detonation is pressure changes are finished. In the OH* time profiles, the light
verified between P1 and P2 at 155 mm. The start and end times from emission begins very close to the nozzle entrance, due to the higher
pressure curves of the P5 transducer (Figs. 4 and 5) are those applied detector sensitivity in the UV range.
for the integrations under pressure curves to determine the specific These first peaks in the emission profiles are related to the shock on
impulses, as shown in Fig. 6. the expansion region under vacuum (10 Pa) and present too high a
The same procedure as that applied by Zhang et al [20] was fol- degree of intensity fluctuations, probably due to the pressure peak
lowed in this work. The initial pressure (101.325 kPa) was considered changes. The same is verified in the pressure curves from the P5
for time integration, and when it returns to zero pressure the end of transducer, as reflected shocks (second peak in Fig. 6). The shock sig-
integration is established, in spite of the pressure decay not being nals disappeared in a few experiments with detonation failure.
completed. Eq. (1) shows the impulse for a PDE with ambient pressure As can be verified in Fig. 3, the OH* and H2O* emission profiles are
(Pa ) under vacuum, as is the case in this work. highly dissimilar. The OH* emission follows the entire detonation

264
N.C. Lopes et al. Experimental Thermal and Fluid Science 102 (2019) 261–270

Time at P1 (120 mm) - τDDT expansion, while that of H2O* is predominantly from the thrust surface.
Start time at P5 (1502 mm)
Fig. 6 shows that the pressure change finishes at around 6 ms, where the
End time at P5 (1502 mm)
Start time at P6 (1550 mm) OH* emission is also finished and that of H2O* is close to the maximum
Time of pressure peak at P6 (1550 mm) Φ = 1.2 intensity (Fig. 3). Furthermore, Figs. 4 and 5 also display these differ-
4.2
4.1
ences, in which higher time values with larger fluctuations were found
4.0
from the H2O* emission profiles, except the start times.
3.9
The third peak from the OH* emission profiles (Fig. 3) and the start
3.8
of the second peak from the H2O* emission profiles (not shown here)
3.7
determine the cycle times (tcycle). The cycle times from OH* emission
profiles (Figs. 4 and 5) were in agreement with the end times from the
Time (ms)

3.6
3.5
pressure curves at P5 (Figs. 4 and 5) and they were determined with
1.75 higher accuracy.
1.70 The emission data displayed in Figs. 4 and 5 were also applied to
1.65 establish the time integrations under the emission profiles, and corre-
1.60
lations to specific impulses, as well as to determine the detection con-
ditions for emission image acquisition.
1.55
Fig. 7 shows the detonation velocities and the peak pressures at P5
1.50
(1502 mm) for stoichiometric (Φ = 1.0) and fuel-rich (Φ = 1.2) H2-air
1.45
1 2 3 4 5 6 7 8 9
detonation waves.
Experimental Test In the majority of the experimental tests with stoichiometric con-
ditions, the velocities slightly decrease to approximately ± 2% of the
OH* start time
Time at peak 3 - τcycle
Chapman–Jouguet velocity (DCJ = 1969 m/s) at the nozzle entrance.
H2O* start time
Time at peak 4 Time at peak However, the detonations exhibited varying behaviours. At least two
Total time of peak 3 Total time of peak unusual conditions could be observed in the first set, in which the ve-
Total time of peak 4 Total emission time
Total emission time locities were either kept constant (tests 3 and 10) or slightly accelerated
6.5
(test 9) to the expansion entry. For fuel-rich conditions, there were
6.0 again two unusual experiments, in which the velocities remained nearly
5.5 constant (tests 3 and 8). However, for the greater number of the fuel-
5.0 rich detonation tests, the detonation waves always accelerated in the
4.5
tube and preponderantly decelerated at the expansion beginning, re-
sulting in velocities below (≤3.5%) the Chapman–Jouguet velocity
4.0
(DCJ = 2030 m/s) at the nozzle entry.
Time (ms)

3.5
All experimental tests for both stoichiometric and fuel-rich condi-
1.5 tions resulted in overpressures at the end of the detonation tube. The
high pressures measured were in the mean 51% for stoichiometric de-
1.0 tonations and 42% for fuel-rich detonations above their
Chapman–Jouguet pressures.
0.5 The detonation data, presented in Figs. 4, 5 and 7, are highly re-
Φ = 1.2 peatable and reproducible, and can represent the studied conditions.
0.0 Velocity deviations of the mean were found to be less than 2% for both
1 2 3 4 5 6 7 8 9 the stoichiometric and fuel-rich H2-air detonations. Similarly, devia-
Experimental Test
tions of approximately 2% and 1% were observed for the times at P5
Fig. 5. Times from pressure curves and times from OH* and H2O* time profiles and P6 in the stoichiometric and fuel-rich conditions, respectively.
for each experimental test of the fuel-rich H2-air detonation. However, slightly higher deviations of the mean were determined for
detonation initiation times, namely ca. 2.5% for stoichiometric and 4%
for fuel-rich detonations. The greatest deviations were those of the
3000 pressure at the end of the detonation tube, where a jump condition
Pressure signal from P5 occurs, with 5% for stoichiometric and 8% for fuel-rich conditions.
2500
Despite the low deviations of the mean found for experimental tests,
in fact, the greater number of the studied detonations exhibited over-
2000
driven propagation and overpressures in the detonation tube, as ob-
served previously. The high pressures resulted in short expansion times
Pressure (kPa)

1500
and, consequently, reduced performance. The specific impulses were
determined according to Eqs. (2) and (3) as the nozzle exit area (m2),
1000
reaction mixture mass (kg), fuel mass (kg) and gravity acceleration (m/
500
s2) were known. The impulses (N s) were established by means of Eq.
(1) and integration under pressure curves at the end of the detonation
0 tube.
IA
-500 Isp =
mmixture g (2)
-3 -3 -3 -3 -3 -3
4.5x10 5.0x10 5.5x10 6.0x10 6.5x10 7.0x10
IA
Time (s) Ispf =
mH2 g (3)
Fig. 6. Pressure curve at the end of the detonation tube (P5 at 1502 mm) for
stoichiometric H2-air detonation. Furthermore, the specific impulses for the ideal PDE were calculated
using the analytical model from Wintenberger and Cooper [56,57].

265
N.C. Lopes et al. Experimental Thermal and Fluid Science 102 (2019) 261–270

2200 5500
2150
--- DCJ = 1969 m/s Pressure 5250
5000
4750
2100 4500
4250
2050 4000
2000 3750
3500
1950 3250
3000
1900 2750
2500

Pressure (kPa)
Velocity (m/s)

1850 2250
2000
1250 1900

1200
1800
1150
1700
1100
1600
1050
P1-P2 P3-P4 P4-P5 P5-P6 Φ = 1.0
1000 1500
1 2 3 4 5 6 7 8 9 10 11
Experimental Test
2200 5500
--- DCJ = 2030 m/s Pressure 5250
2150 5000
4750
2100 4500
4250
2050 4000
3750
2000 3500
3250
1950 3000
2750
1900 2500

Pressure (kPa)
2250
Velocity (m/s)

1850 2000
1750
1250 1700

1200

1150
1600
1100

1050
P1-P2 P3-P4 P4-P5 P5-P6 Φ = 1.2
1000 1500
1 2 3 4 5 6 7 8 9
Experimental Test
Fig. 7. Velocities from P1–P2 position (155 mm), P3–P4 position (1330 mm), P4–P5 position (1434 mm) and P5–P6 (1526 mm) and peak pressures at P5 for each
experimental test of the H2-air detonations. Some velocities could not be determined due to the high amount of electromagnetic noise from the spark discharge. The
error bars are associated with pressure transducer time response and the distance between each pair of pressure transducers. Chapman–Jouguet pressures (PCJ) for
stoichiometric (Φ = 1.0) and fuel-rich (Φ = 1.2) detonations are 1580.7 kPa and 1590.8 kPa, respectively.

Table 1 DDT time (tDDT) plus cycle time (tcycle), while maximum thrust (Tmax) is
Parameters for H2-air detonation cycle performance. defined as follows.
Isp (s) Ispf (s) tcycle (ms) fmax (Hz) Tmax (N)
Tmax = mmixture gfmax Isp (4)
Stoichiometric H2-air detonations (Φ = 1.0)
Calculated 256 8979 10.75 93 309 The specific impulses, cycle times, maximum frequency and max-
Experimental 109 ± 2 3826 ± 74 5.25 ± 0.05 135 ± 7 181 ± 14 imum thrust are displayed in Table 1.
Fuel-rich H2-air detonations (Φ = 1.2) The experimental data values were significantly lower than those
Calculated 264 7745 10.45 95 311 calculated for the ideal PDE, as shown in Table 1. This is mainly as a
Experimental 112 ± 3 3286 ± 76 4.19 ± 0.05 174 ± 11 240 ± 33 result of the overpressure propagating detonations, associated with
detonation initiation with a large turbulence. For improved perfor-
mance, the length and BR of the Shchelkin should be optimized for the
Cycle times, maximum frequency and thrust were also calculated H2-air mixtures and geometry of the detonation chamber, or direct
[56,58] and determined via experiments. Experimental DDT times were detonation initiation should be applied (tDDT = 0). However, despite
obtained by means of the first transducer’s pressure curves and cycle the inferior performance obtained, the established frequencies were
times were established from the OH* emission profiles, as reported within the possible operating frequency range for a valveless PDE [59].
previously. Maximum frequency (fmax) is defined by the inverse of the Specific impulses versus integrated areas from the OH* and H2O*

266
N.C. Lopes et al. Experimental Thermal and Fluid Science 102 (2019) 261–270

15000 OH* peak area 15000 36000


Sum of OH* peaks
14000 14000

Sum of OH* peaks (u.a. kg )


34000

-1

-1
OH* Peak Area (u.a. kg )
-1

OH* Total Area / u.a. kg


13000 13000
32000

12000 12000
30000

11000 11000
28000

10000 10000
26000

9000 9000
3100 3200 3300 3400 3500 3600 3700 3800 3900 3100 3200 3300 3400 3500 3600 3700 3800 3900

Ispf / s Ispf / s

950 90000

900
80000
H2O* Peak Area (u.a kg )

H2O* Total Area (u.a kg )


-1

-1
850
70000

800

60000
750

50000
700

650 40000
108 110 112 114 116 118 3200 3300 3400 3500 3600 3700 3800 3900 4000
Isp / s Ispf / s

Fig. 8. Correlations between emission time profiles and specific impulses based on fuel (Ispf) and mixture (Isp). The OH* emission vs. Ispf graphs are shown at the top,
and the H2O* emission vs. Isp and Ispf graphs at the bottom.

time profiles were plotted, to establish possible correlations. The areas 90% of the total H2O* areas to the Ispf values, due to the hydrogen mass
under peaks 3 and 4, as well as areas under the total time profile, ne- being directly proportional to the amount of H2O* production.
glecting the first peak from the OH* emission profiles (Fig. 3), per fuel These results could aid the development of optoelectronic sensors
mass, were considered. For the H2O* emission profiles, the areas under for detonation engines, with the correlations as in this work applied for
the main peak, and those under total emission, per fuel mass and per calibration. Similarly, emission probes are usually employed in the
reaction mixture mass were analysed. automotive industry for injection control and engine optimization.
Fig. 8 shows that the OH* emission correlates only with the specific The same diagnostic method presented here could furthermore be
impulse based on fuel (Ispf), while the H2O* emission correlates with extended to achieve spontaneous emission acquisition through high-
both specific impulses. speed cameras.
Although H2O* emission areas correlate with both specific impulses, Single-shot OH* emission images were captured to identify the de-
it is important to note that H2O* emission times from detonation were tonation occurrence in the exhaust plume, their dimensions and pro-
always with higher fluctuations than OH* emission times (Figs. 4 and pagation regime. OH* images were acquired by collecting a tiny portion
5). of the large peak 4, to avoid loss of synchronization time and reduce
A non-linear exponential growth was fitted to the OH* areas versus intensity fluctuations for acquisition. Delay times adjusted to the ICCD
Ispf data with correlations of 94% when applying the sum of peaks, 90% camera were determined by means of time emission data (Figs. 4 and
when using only peak 4 (76% of the sum), and 95% when the total 5), while exposure times were established using spectral response
emission area was considered. The OH* emission has a strong connec- curves of the optical components, as well as detectors of the mono-
tion to the detonation front’s structure [60,61,62], and, it should, chromator and ICCD camera associated with the OH* emission in-
therefore, be possible to evaluate the detonation cycle performance tensity, recorded as a function of time.
through its radiative emission. The OH* areas are related to the radical Fig. 9 presents the OH* images at the nozzle exit for stoichiometric
amount per kilogram of fuel [63], and the higher the OH* amount, the and fuel-rich H2-air detonations.
higher Ispf and the lower the performance. The OH* emission images exhibit sustained detonations at the
For H2O* emission data, the non-linear exponential decay of the plume, with similar features as those recently observed in the studies on
main peak area and Isp curve was fitted with a correlation of 93%, H2-air detonations in an explosion channel [60,62] and the rotating
probably due to the effect of the dissociation on H2O* emissions for the detonation engine (RDE) [61].
spectral range applied. The lower the dissociation of molecules, the The OH* images reveal multi-headed detonations in the stoichio-
higher the power, as less heat is lost. It follows that there will be a metric H2-air case (Fig. 7a and b), despite two very different images
minor contribution to the continuum emission in the visible range as- being captured, probably due to the images not being at the exact same
sociated with the H2O molecules. In contrast, a linear fit correlates with propagation time. The multi-headed regime is stable detonation

267
N.C. Lopes et al. Experimental Thermal and Fluid Science 102 (2019) 261–270

Fig. 9. OH* emission images from expansion of


the H2-air detonation, where exposure times of
2.5 μs were applied: (a), (b) stoichiometric mix-
tures (Φ = 1.0) with similar Isp, tests 10 and 11
and (c), (d) fuel-rich mixtures (Φ = 1.2), where
the image with the lowest Isp is displayed on the
left and that with the highest Isp on the right
(tests 8 and 9).

propagation, and a height of at least three detonation cells is necessary measurements for an ideal detonation cycle. Velocities and specific
to achieve this propagation regime [60]. impulses from pressure curves, integrated areas from emission time
Fig. 9a presents three regions of higher emission intensity related to profiles and OH* emission images were determined for the performance
the detonation cells, and Fig. 9b displays a narrow region with several characterization and evaluation of the detonation cycle.
centres of high luminosity intensity with smaller sizes. Highly repeatable and reproducible H2-air detonation parameters
For fuel-rich H2-air detonations, the OH* images reveal a two- were obtained for the detonation cycles ignited with a 65 J microsecond
headed detonation propagation, which corresponds to a transition re- discharge, coupled to a Shchelkin of 580 mm length. Deviations of the
gime that exhibits less stable propagation, but is still controlled, and mean were found to be less than 2% for velocity data and less than 4%
stable propagating detonations. The unstable single-headed detonation for detonation initiation times. Analogously, the specific impulses ex-
with one strong transverse wave can be interpreted as the near-limit hibited deviations of less than 2.7%. The pressure values at the end of
propagation of detonation, similar to spinning detonations, where de- the detonation tube were those with higher fluctuations, with ap-
tonation failure may occur [60]. proximately 5% and 8% for the stoichiometric and fuel-rich conditions,
Our results are in accordance with H2-air detonations for the RDE, respectively.
where a consistent multi-headed detonation for the stoichiometric Correlations were determined to exist between both the emitters
condition, and two-headed to occasionally single-headed detonation for and specific impulses, where those with integrated areas under the
the fuel-rich mixture with a higher equivalence ratio (Φ = 1.3), were emission peaks from the thrust surface were more strongly correlated.
observed [61]. The OH* amount per kilogram of fuel exhibited an exponential corre-
Furthermore, the reaction zone widths in these images were de- lation to the Ispf. The higher the OH* amount, the lower the detonation
termined by applying the free IC Measure software. After performing cycle performance because the Ispf is higher, and lower OH* radical
calibration to define the image scale, a 15–20 mm width for stoichio- amounts were produced in the fuel-rich mixtures. Furthermore, the
metric H2-air and an 18–24 mm width for the fuel-rich H2-air detona- H2O* emission peak area per kilogram of mixture exhibited an ex-
tions were established. These dimensions are related to the detonation ponential correlation to the Isp. The lower the H2O* emission area, the
cell lengths, which were calculated (λell5 = 0.6 Lcell) as 16.7 and higher the power due to the occurrence of a lower dissociation of H2O*
16.3 mm for the H2-air stoichiometric and fuel-rich (Φ = 1.2) detona- molecules and their resultant smaller contribution to the continuum
tions, respectively. emission.
The OH* imaging results characterize the detonation process at the The OH* images validate the established correlations because they
nozzle exit and corroborate the optical diagnostic method employed. indicate detonations with stable propagations at the nozzle exit for both
studied conditions. For the stoichiometric H2-air detonations, multi-
headed propagating detonations were observed, while two-headed de-
4. Conclusions tonations are shown in the images for the fuel-rich mixtures.
Therefore, an ideal detonation cycle with ignition and geometry
In this study, we carried out pressure and spontaneous emission features, as in this work, can achieve reliable and stable propagations
with H2-air mixtures from stoichiometric to the equivalence ratio of 1.2.
5 However, combustible mixtures with more hydrogen could lead to
λ is the detonation cell width.

268
N.C. Lopes et al. Experimental Thermal and Fluid Science 102 (2019) 261–270

detonation failure, and the Shchelkin dimensions should be optimized 586–596, https://doi.org/10.2514/1.B34013.
for improving performance. [27] M. Chou, F.E. Fendell, H.W. Behrens, Theoretical and experimental studies of laser-
initiated detonation waves for supersonic combustion, Proc. SPIE 1862, Laser Appl.
A new method based on spontaneous emissions to evaluate deto- Combust. Combust. Diagnostics, 1993, pp. 45–58, , https://doi.org/10.1117/12.
nation cycles was developed, and this optical approach can be applied 2252608.
to optoelectronic sensor calibration for monitoring the performance of [28] A. Starikovskiy, N. Aleksandrov, A. Rakitin, Plasma-assisted ignition and deflagra-
tion-to-detonation transition, Philos. Trans. A. Math. Phys. Eng. Sci. 370 (2012)
detonation engines. 740–773, https://doi.org/10.1098/rsta.2011.0344.
[29] F. Wang, A. Kuthi, M.A. Gurdensen, Technology for transient plasma ignition, Proc.
Acknowledgements 43rd AIAA Aerosp. Sci. Meet. Exhib. 2005, pp. 1–9.
[30] A.E. Rakitin, A.Y. Starikovskii, Mechanisms of deflagration-to-detonation transition
under initiation by high-voltage nanosecond discharges, Combust. Flame 155
The authors thank Diagnostic Techniques Subdivision for the ICCD (2008) 343–355, https://doi.org/10.1016/j.combustflame.2008.05.019.
camera and Experimental Hypersonics Subdivision for the delay gen- [31] C.S.T. Marques, A.C. Oliveira, F. Dovichi Filho, W.C. Ferraz, J.B. Chanes Jr., Single-
shot pulsed detonation device for pde combustion simulation, Proc. 13th Brazilian
erator. This work was supported by FAPESP (Fundação de Amparo à
Congr. Therm. Sci. Enginering, 2010, pp. 1–10.
Pesquisa do Estado de São Paulo) through projects 2008/10548-5 and [32] F.C.F.D. Bidia, C.S.T. Marques, A.P. Pimenta, J.B.R. Santos, Diagnóstico da estru-
2013/26708-0, and by CNPq (Conselho Nacional de Pesquisa) through tura de frentes de detonação por meio de folhas com fuligem, Ciência Eng. 23
project 471052/2012-4. The authors also thank CNPq (145070/2015-8 (2014) 91–100.
[33] Z. Wang, Y. Zhang, J. Huang, Z. Liang, L. Zheng, J. Lu, Ignition method effect on
and 123238/2016-1) and COMAER PROHIPER for fellowships. detonation initiation characteristics in a pulse detonation engine, Appl. Therm. Eng.
93 (2016) 1–7.
References [34] P.K. Panicker, J.-M. Li, F.K. Lu, D.R. Wilson, Development of a compact liquid
fueled pulsed detonation engine with predetonator, Proc. 45th AIAA Aerosp. Sci.
Meet. Exhib. 2007, pp. 1–13.
[1] P.M. Rubins, R.C. Bauer, Review of shock-induced supersonic combustion research [35] S.M. Frolov, Liquid-fueled, air-breathing pulse detonation engine demonstrator:
and hypersonic applications, J. Propul. Power 10 (1994) 593–601, https://doi.org/ operation principles and performance, J. Propul. Power 22 (2006) 1162–1169,
10.2514/3.23768. https://doi.org/10.2514/1.17968.
[2] J.P. Sislian, Detonation wave ramjet, in: S.N.B. Murthy, M.E.T. Curran (Eds.), [36] R. Driscoll, W. Stoddard, A.S. George, E. Gutmark, Shock transfer and shock-in-
Scramjet Propuls., Progress in Astronautics and Aeronautics, 2000, pp. 823–889. itiated detonation in a dual pulse detonation engine/crossover system, AIAA J. 53
[3] F.K. Lu, Prospects for detonations in propulsion, Proc. 9th Int. Symp. Exp. Comput. (2015), https://doi.org/10.2514/1.J053027.
Aerothermodyn. Intern. Flows, 2009, pp. 1–10. [37] S.M. Frolov, V.Y. Basevich, V.S. Aksenov, S.A. Polikhov, Optimization study of
[4] P. Wolański, Detonative propulsion, Proc. Combust. Inst. 34 (2013) 125–158, spray detonation initiation by electric discharges, Shock Waves 14 (2005) 175–186,
https://doi.org/10.1016/j.proci.2012.10.005. https://doi.org/10.1007/s00193-005-0263-8.
[5] J.A. Nicholls, H.R. Wilkinson, R.B. Morrison, Intermittent detonation as a thrust- [38] S.M. Frolov, V.Y. Basevich, V.S. Aksenov, S.A. Polikhov, Detonation initiation by
producing mechanism, J. Jet Propul. 27 (1957) 534–541, https://doi.org/10.2514/ controlled triggering of electric discharges, J. Propul. Power 19 (2003) 573–580,
8.12851. https://doi.org/10.2514/1.5969.
[6] K. Kailasanath, Review of propulsion applications of detonation waves, AIAA J. 38 [39] D.H. Lieberman, J.E. Shepherd, Detonation initiation by a hot turblent jet for use in
(2000) 1698–1707. pulse detonation engines, Proc. 38th AIAA/ASME/SAE/ASEE Jt. Propuls. Conf.
[7] K. Kailasanath, Recent developments in the research on pulse detonation engines, Exhib. 2002, pp. 1–9, , https://doi.org/10.2514/6.2002-3909.
AIAA J. 41 (2003) 145–149. [40] H. Shimada, Y. Kenmoku, H. Sato, A.K. Hayashi, A new ignition system for pulse
[8] https://www.flightglobal.com/news/articles/us-afrl-proves-pulse-detonation- detonation engine hidetoshi, Proc. 42nd AIAA Aerosp. Sci. Meet. Exhib. 2004, pp.
engine-can-power-aircraft-222008/ (accessed March 15, 2017). 1–9, , https://doi.org/10.2514/6.2004-308.
[9] L. Butler, L.W. Dunbar, J.E. Johnson, US pat 6442930 B1, 2001. [41] R. Sorin, R. Zitoun, D. Desbordes, Optimization of the deflagration to detonation
[10] A. Rasheed, A.J. Dean, US Pat 7784265 B2, 2010 (2007:874974). transition, Shock Waves 15 (2006) 137–145, https://doi.org/10.1007/s00193-006-
[11] R.B. Edelman, H.R. Lander, J.H. Hunt, US pat 20080087010, 2008 (CA 0007-4.
148:452583). [42] M. Cooper, S. Jackson, J. Austin, E. Wintenberger, J.E. Shepherd, Direct experi-
[12] P.H. Shimo, M. Snyder, WO Pat 2012/092285, WO/2012/092285, 2011. mental impulse measurements for detonations and deflagrations, J. Propul. Power
[13] F. Giuliani, A. Lang, M. Irannezhad, A. Lundbladh, M. Dynamics, C. Systems, 18 (2002) 1–41, https://doi.org/10.2514/2.6052.
V.A. Corporation, Pulse detonation as an option for future innovative gas turbine [43] K. Asato, T. Miyassaka, Y. Watanabe, K. Tanabashi, Combined effects of vortex flow
combustion technologies, Proc. 27th Int. Congr. Aeronaut. Sci. 2010, pp. 1–10. and the shchelkin spiral dimensions characteristics of deflagration-to-detonation
[14] https://www.technologyreview.com/s/507421/exploding-engine-could-reduce- transition, Shock Waves 23 (2013) 325–335.
fuel-consumption/ (accessed March 16, 2017). [44] E. Wintenberger, J.E. Shepherd, Thermodynamic cycle analysis for propagating
[15] F.K. Lu, D.R. Wilson, Some perspectives on pulse detonation propulsion systems, detonations, J. Propul. Power 22 (2006) 694–698.
Proc. 24th Int. Symp. Shock Waves, 2004, pp. 1–14. [45] J.R. Gord, M.S. Brown, T.R. Meyer, Optical diagnostics for characterizing advanced
[16] P.R. Panicker, D.R. Wilson, F.K. Lu, Operational issues affecting the practical im- combustors and pulsed-detonation engines, Proc. 22nd AIAA Aerodyn. Meas.
plementation of pulse detonation engines, Proc. 14th AIAA/AHI Sp. Planes Technol. Gr. Test. Conf. 2002, pp. 1–14.
Hypersonic Syst. Technol. Conf. 2006, pp. 1–18. [46] K. Matsuoka, M. Esumi, K. Bryan, J. Kasahara, A. Matsuo, I. Funaki, Optical and
[17] S. Lee, J. Watts, S. Saretto, S. Pal, C. Conrad, R. Woodward, R. Santoro, Deflagration thrust measurement of a pulse detonation combustor with a coaxial rotary valve,
to detonation transition processes by turbulence-generating obstacles in pulse de- Combust. Flame 159 (2012) 1321–1338, https://doi.org/10.1016/j.combustflame.
tonation engines, J. Propul. Power. 20 (2004). 2011.10.001.
[18] J.A. Dean, I.B. Schild, Influence of spark energy, spark number, and flow velocity on [47] V.E. Tangirala, A.J. Dean, P.F. Pinard, B. Varatharajan, Investigations of cycle
detonation initiation in a hydrocarbon-fueled PDE, Proc. 41st AIAA/ASME/SAE/ processes in a pulsed detonation engine operating on fuel – air mixtures, Proc.
ASEE Jt. Propuls. Conf. Exhib. (2005) 1–7. Combust. Inst. 30 (2005) 2817–2824, https://doi.org/10.1016/j.proci.2004.08.
[19] J. Kasahara, K. Takazawa, T. Arai, Y.U. Tanahashi, S. Chiba, A. Matsuo, 208.
Experimental investigations of momentum and heat transfer in pulse detonation [48] F.R. Schauer, C.L. Miser, K.C. Tucker, R.P. Bradley, J.L. Hoke, Detonation initiation
engines, Proc. Combust. Inst. 29 (2002) 2847–2854. of hydrocarbon-air mixtures in a pulsed detonation engine, Proc. 43rd AIAA Aerosp.
[20] F.Y. Zhang, T. Fujiwara, T. Nyiasaka, E. Nakayama, T. Hattori, N. Azuma, Sci. Meet. (2005) 1–10.
S. Yoshida, S. Tamugi, Experimental study of key issues on pulse detonation engine [49] J.E. Shepherd, F. Pintgen, J.M. Austin, C.A. Eckett, The structure of the detonation
development, Trans. Japan Soc. Aero. Sp. Sci. 45 (2003) 243–248. front in gases, Proc. 40th AIAA Aerosp. Sci. Meet. Exhib. 2002, pp. 1–13.
[21] C.M. Brophy, R.K. Hanson, Fuel distribution effects on pulse detonation engine [50] C. Xiong, C.J. Yan, H. Qiu, W. Fan, Multicycle detonation investigation by emission-
operation and performance, J. Propul. Power. 22 (2006) 1155–1161, https://doi. absorption-based temperature diagnostics, Combust. Sci. Technol. 183 (2011)
org/10.2514/1.18713. 62–74, https://doi.org/10.1080/00102202.2010.502551.
[22] Z.C. Owens, R.K. Hanson, Single-cycle unsteady nozzle phenomena in pulse deto- [51] X. Wan, Z. Zhang, X. Leng, B. Deng, Three dimensional measurements of engine
nation engines, J. Propul. Power. 23 (2007) 325–337, https://doi.org/10.2514/1. plumes with four-channel single spectral tomography three dimensional measure-
22415. ments of engine plumes with four-channel single spectral tomography, J. Appl.
[23] A.D. Cutler, High-frequency pulsed combustion actuator experiments, AIAA J. 49 Phys. 108 2010, pp. 73107-1–73107-6, , https://doi.org/10.1063/1.3490234.
(2011) 1943–1950, https://doi.org/10.2514/1.J050876. [52] D.W. Mattison, C.M. Brophy, S.T. Sanders, L. Ma, K.M. Hinckley, J.B. Jeffries,
[24] T. Ombrello, C. Carter, J. McCall, F. Schauer, C. Tam, A. Naples, J. Hoke, K.Y. Hsu, R.K. Hanson, Pulse detonation engine characterization and control using tunable
Enhanced mixing in supersonic flow using a pulse detonation combustor, Proc. 50th diode-laser sensors introduction, J. Propul. Power 19 (2003) 568–572.
AIAA Aerosp. Sci. Meet. Incl. New Horizons Forum Aerosp. Expo. 2012, pp. 1–15. [53] R.K. Hanson, Applications of quantitative laser sensors to kinetics, propulsion and
[25] K.P. Rouser, P.I. King, F.R. Schauer, R. Sondergaard, J.L. Hoke, L.P. Goss, Time- practical energy systems, Proc. Combust. Inst. 33 (2011) 1–40, https://doi.org/10.
resolved flow properties in a turbine driven by pulsed detonations, J. Propul. 1016/j.proci.2010.09.007.
Power. 30 (2014) 1528–1536, https://doi.org/10.2514/1.B34966. [54] D.W. Mikolaitis, C. Segal, A. Chandy, Ignition delay for jet propellant 10/air and jet
[26] A. Rasheed, A.H. Furman, A.J. Dean, Experimental investigations of the perfor- propellant 10/high-energy density fuel/air mixtures, J. Propul. Power 19 (2003),
mance of a multitube pulse detonation turbine system, J. Propul. Power. 27 (2011) https://doi.org/10.2514/2.6147.

269
N.C. Lopes et al. Experimental Thermal and Fluid Science 102 (2019) 261–270

[55] R.P. Lindstedt, H.J. Michels, Deflagration to detonation transitions and strong de? hydrogen–air mixtures with transverse concentration gradients, Shock Waves 26
agrations in alkane and alkene air mixtures, Combust. Flame 76 (1989) 169–181. (2016) 181–192, https://doi.org/10.1007/s00193-015-0598-8.
[56] E. Wintenberger, J.M. Austin, M. Cooper, S. Jackson, J.E. Shepherd, Analytical [61] B.A. Rankin, D.R. Richardson, A.W. Caswell, A.G. Naples, J.L. Hoke, F.R. Schauer,
model for the impulse of single-cycle pulse detonation tube, J. Propul. Power 19 Chemiluminescence imaging of an optically accessible non-premixed rotating de-
(2003) 22–38, https://doi.org/10.2514/2.6099. tonation engine, Combust. Flame 176 (2016) 12–22, https://doi.org/10.1016/j.
[57] M. Cooper, J.E. Shepherd, The effect of transient nozzle flow on detonation tube combustflame.2016.09.020.
impulse, Proc. 40th AIAA/ASME/SAE/ASEE Jt. Propuls. Conf. Exhib. 2004, pp. [62] L.R. Boeck, F.M. Berger, J. Hasslberger, F. Ettner, T. Sattelmayer, Macroscopic
1–22, , https://doi.org/10.2514/6.2004-3914. structure of fast deflagrations and detonations in hydrogen-air mixtures with con-
[58] T. Endo, T. Fujiwara, A simplified analysis on a pulse detonation engine model, centration gradients, Proc. 24th ICDERS, 2013, pp. 1–6.
Tran. Jap. Soc. Aero. Sp. Sci. 44 (2002) 217–222. [63] C.S.T. Marques, L.H. Benvenutti, C.A. Bertran, Experimental study of OH*, CHO*,
[59] K. Matsuoka, T. Mukai, T. Endo, Liquid-purge method in high frequency valveless CH* and C2* radicals in C2H2/O2 and C2H2/O2/Ar flames in a closed chamber,
pulse detonation engine, Proc. 24th ICDERS, 2013, pp. 1–6. Combust. Sci. Technol. 167 (2001) 113–129.
[60] L.R. Boeck, F.M. Berger, J. Hasslberger, T. Sattelmayer, Detonation propagation in

270

You might also like