You are on page 1of 67

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/224997457

Evaluation of full engine scramjet technology

Article · July 2004


Source: DLR

CITATIONS READS

8 3,502

2 authors:

Anthony D. Gardner Klaus Hannemann


German Aerospace Center (DLR) German Aerospace Center (DLR)
163 PUBLICATIONS   1,587 CITATIONS    427 PUBLICATIONS   2,698 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

FAST Rescue View project

Combustion Chamber Modelling View project

All content following this page was uploaded by Anthony D. Gardner on 28 May 2014.

The user has requested enhancement of the downloaded file.


   

Berichts.-Nr.: 
DLR-IB 224-2004 A 11 

Verfasser: 
A.D. Gardner, K. Hannemann 
 

Titel: 
Evaluation of Full-Engine
Scramjet Technology

   

Datum:  2.12.2004 
Auftraggeber:  ESA-ESTEC 
Auftrags-Nr.:  TRP 17001/02/NL/MV 
Vorgesehen für:   
Zugänglichkeitsstufe:  1 

  Der Bericht umfaßt: 


64  Seiten 
5  Tabellen 
68  Bilder 
19  Literaturstellen 

Vervielfältigung und Weitergabe dieser Unterlagen sowie Mitteilung ihres Inhalts an Dritte,  
auch auszugsweise, nur mit Genehmigung           X  des DLR        X  des Auftraggebers. 

DLR
Institut für Aerodynamik und Strömungstechnik
Abteilung Raumfahrzeuge
Bunsenstraße 10
37073 Göttingen
DLR

Evaluation of Full-Engine  
Scramjet Technology 

Summary:

The  first  phase  of  the  HyShot  supersonic  combustion  ramjet  (scramjet)  flight 
experiment  program  of  The  University  of  Queensland  in  Australia  was  designed  to 
provide  benchmark  data  on  supersonic  combustion  for  a  flight  Mach  number  of 
approximately  M=8.  The  second  flight  of  the  HyShot  program,  performed  on  July 
30th  2002,  was  successful  and  supersonic  combustion  was  observed  along  the 
specified trajectory range. 

The  operating  range  of  the  High  Enthalpy  Shock  Tunnel  Göttingen  (HEG)  of  the 
German  Aerospace  Center  (DLR)  was  recently  extended.  The  facility  now  has  the 
capability  of  testing  a  complete  scramjet  engine  with  internal  combustion  and 
external aerodynamics at M=7.8 flight conditions in altitudes of about 30km. A post-
flight  analysis  of  the  HyShot  flight  experiment  was  performed  under  an  ESA-ESTEC 
TRP contract using an operational scramjet wind tunnel model with a geometry which 
is identical to that of the flight configuration. 

DEUTSCHES ZENTRUM FÜR LUFT-


UND RAUMFAHRT E.V.

Institut für Aerodynamik und Strömungstechnik


Abteilung Raumfahrzeuge 

Institutsleiter:  Verfasser: 
   
   
   
   
(Prof. Dr. A. Dillmann)  (A.D. Gardner)  
   
Abteilungsleiter:  Verfasser: 
   
   
   
   
(Dr. K. Hannemann)  (Dr. K. Hannemann) 
   
Datum:  2.12.2004  Abteilung:  Bericht: 
Bearbeitet:  K. Hannemann  Raumfahrzeuge  224-2004 A 11 
Contents
Table of contents 1
List of Symbols 2
Introduction 3
1 HYSHOT Flight Experiment 5
1.1 Mission Profile Overview 5
1.2 Design Overview 7
1.3 Instrumentation Details 10
1.4 Fuel System 12
2 HEG Facility and Ground Test Model 13
2.1 The High Enthalpy Shock Tunnel Göttingen (HEG) 13
2.2 HyShot Ground Test Model Design 18
2.3 Instrumentation of Ground Test Model 20
3 HEG Ground Testing 21
3.1 Methods for the Analysis of Results 27
3.2 Results and Discussion 34
4 Conclusions 38
5 Comments and Recommendations for Future Activities 39
References 42
A1 Experimental Results for each Individual Run 43
A1.1 Run 662 43
A1.2 Run 663 45
A1.3 Run 664 47
A1.4 Run 666 49
A1.5 Run 667 51
A1.6 Run 668 53
A1.7 Run 670 55
A1.8 Run 671 59
A1.9 Run 672 63

1
List of Symbols
α Model angle of attack. Pitch.
Cp Specific heat at constant pressure.
γ Ratio of specific heats
H0 Stagnation enthalpy.
K Fuel system calibration constant.
KPt2 Constant to provide mean Pitot pressure from the Pitot pressure measured by the permanent
probe.
mD Mass flow rate.
M Mach number.
Mwt Molecular weight of species.
φ Equivalence ratio assuming one-equation combustion: 2H2+O2H20
P0 Total pressure.
Pstat Static pressure.
Pt2 Pitot pressure.
qD Heat flux.
q00 Stagnation point heat flux on the HEG permanent probe
ρ Density.
R Gas constant.
Re Reynolds number.
τi Huber initiation time.
τr Huber reaction time
Tvib Vibrational temperature of a species (When different to the translational temperature).
Tstat Static temperature.
V Volume
vx,vy Velocity in x, y direction.

2
Introduction
The operating range of the High Enthalpy Shock Tunnel Göttingen (HEG) of the German Aerospace
Centre (DLR) was recently extended. The facility has now the capability of testing a complete
supersonic combustion ramjet (scramjet) engine with internal combustion and external aerodynamics.
A post flight analysis of the HyShot combustion flight experiment was performed using an operational
scramjet wind tunnel model identical to the flight configuration. This test campaign included detailed
wall pressure and heat transfer measurements for M=7.8 flight conditions in an altitude range between
27 km and 33 km.

The HyShot Flight Program of The University of Queensland in Australia was designed to provide
supersonic combustion flight data for flight Mach numbers in excess of 7.5. Two such flights (HyShot
I and II) have been undertaken from the Woomera Prohibited Area in Australia. The first flight was
undertaken on October 30th 2001 and the second on July 30th 2002. While the first flight was
unsuccessful at delivering the experiment to its planned trajectory, the second flight was successful
and supersonic combustion was observed.

The present investigation represents the first ground based testing of the HyShot flight geometry and
subsequent comparison of the obtained data with that resulting from the flight experiment. In a further
step it is intended to compare the data measured in HEG with data obtained on the same model in the
T4 shock tunnel of the University of Queensland. The tests in T4 are currently performed. The present
experiments have allowed to acquire the necessary experience for scramjet testing in the HEG and to
gain confidence in the wind tunnel set-up, the measurement techniques and the post-processing
techniques for on-ground scramjet simulations.

A mirrored test-model was used in HEG such that two evaluations could be made during one run. The
configuration of each side was identical to the Hyshot geometry. The Hyshot flight experiment also
consisted of a mirrored model using one side for fuel-on and the opposite side for fuel-off
measurements. Due to the fact that during flight the Hyshot flight experiment was spinning at a rate
between 5Hz and 7Hz with an angle of attack between 3o and 6o, measurements were always taken in
the fuel-on and fuel-off side at the same relative position (at maximum angle of attack). Therefore, the
wind tunnel model was designed in such a way that both sides were at the same angle of attack in
order to allow direct comparison with the flight data. The wind tunnel model was equipped with
pressure gauges (as during flight) and additionally with heat transfer gauges.

The pressure and heat transfer measurements in the Hyshot wind tunnel model were related to flight
altitudes between 37 km and 23 km. Though the original work package description required
measurement at five altitude points (see Table 1), after investigation of two test conditions,
corresponding to flight altitudes of 32.5 km and 27.1 km, it was decided in consultation with the
technical officer of ESA to change the test matrix to further investigate the sensitivity of the engine to
flow and geometry parameters only at these two altitudes. The raw data obtained in the HEG was post-
processed and analysed and the measurements were compared with the Hyshot II flight data.

3
Nominal Nozzle Stagnation Nozzle Nozzle exit Nozzle Nozzle Nozzle exit
flight stagnation Enthalpy reservoir temperature static exit exit density
altitude pressure [MJ/kg] temperature [K] pressure velocity [kg/m3]
[km] [MPa] [K] [Pa] [m/s]
37 3.44 3.13 2874 240 420 2399 0.0061
33 6.55 3.02 2762 229 759 2356 0.0116
29 12.10 2.96 2700 223 1339 2334 0.0211
25 22.50 2.96 2680 222 2415 2335 0.0383
23 30.73 2.97 2682 223 3267 2338 0.0517

Table1:Original design of nominal HEG test conditions.

This report provides discussions of:

• Scramjet wind tunnel model design,


• Description of HEG fuel system,
• Instrumentation of the wind tunnel model and
• Methods and data used for the calibration of the instrumentation,
• Experimental results,
• Analysis of results, comparison with HyShot II flight data, CFD and analytical computations.

This work is undertaken under the aegis of an ESTEC TRP, with contract number 17001/02/NL/MV.
The technical officer in charge was Johan Steelant of ESTEC. On the DLR side, the project was led by
Klaus Hannemann.

The HyShot wind tunnel model was constructed and provided by Allan Paull from The University of
Queensland, Australia. The support of Allan Paull is highly acknowledged here.

4
1 HYSHOT Flight Experiment
The HyShot supersonic combustion flight experiment, successfully performed in July 2002, was led
by The University of Queensland in Australia with contributions from a number of international
partners (Paull et al. (2002)). The HyShot program, which will continue on with more complex
scramjet flow paths in the near future, is an experimental flight program designed to provide
supersonic combustion flight data for flight Mach numbers in excess of 7.5. In the subsequent
sections, an overview of the HyShot I and II flights given.

1.1 Mission Profile Overview


The first two HyShot flights were powered with a two-stage Terrier-Orion Mk70 sounding rocket,
which were launched from Woomera in Australia. After burn out of the second stage, the Orion motor
remained attached to the scramjet module to ensure aerodynamic stability. After reaching the apogee
in an altitude of approximately 315 km, the vehicle was turned and accelerated purely by the
gravitational force during the descent phase (see Fig. 1.1.1.). At a flight Mach number of
approximately M=7.6 and at altitudes from approximately 35 km down to 23 km hydrogen supersonic
combustion data could be successfully collected.

300

Apogee
(278SEC, 314KM, M2.7) Stop Attitude Control Maneuver
(460SEC, 167KM, M6.6)
Start Attitude Control Maneuver

ALTITUDE
[KM] 200
Nosecone Eject
(64SEC, 106KM, M7.1)

Re-enter Atmosphere
(497SEC, 100KM, M7.7)

Stage Separation
(9SEC, 6.5KM, M3.4)

Terrier Burnout
(5.9SEC, 3.4KM, M3.7) 100
Start Experiment
Orion Burnout (527SEC, 35KM, M7.6)
(40SEC, 56KM, M7.4)

Orion Ignition Stop Experiment


(15SEC,12.3KM, M2.8) (533SEC, 23KM, M7.4)

0 100 200 300 400

Terrier Ignition RANGE Impact


(0SEC, 0KM, M0) [KM] (562SEC, 0KM, M0.7)

Figure 1.1.1: Nominal flight profile of HyShot I and II.

As the combustion process in the scramjet is dependant on the ambient pressure, a highly parabolic
trajectory with a near vertical decent during the test time was chosen. This provided supersonic
combustion data with intake pressures that nominally ranged between 30kPa and 180kPa. In addition
to providing an envelope of conditions, a vertical trajectory is also more cost efficient. There are less
structural and environmental difficulties that must be overcome in a highly parabolic flight, because
the heating and dynamic loads are lower than would be experienced if a more level flight was
5
undertaken. This is primarily because in the flat trajectory the energy is provided to the system in the
lower atmosphere, where the density is the greatest. In contrast, in the HyShot profile the rocket
motors provide the majority of the energy in the upper atmosphere, where the density is low. As a
result, the payload does not require active cooling, and is therefore smaller, lighter and less complex.
In addition, the overall drag is lower. Both of which contribute to reducing the size of the rocket
motors. As a result, the program is significantly more cost efficient because relatively small motors
can be used to achieve the same speeds.

With the payload and the second stage re-oriented and pointing downwards, the flight experiment
began when the Pitot pressure rose above a prescribed value (approximately 31kPa), which
corresponded to an altitude of 37km at Mach 7.6. Two combustion chambers were in the model. One
combustion chamber was fuelled, while the other was unfuelled. The experiment was designed to
provide a constant equivalence ratio of 0.3 to the fuelled combustion chamber, until an altitude of
25km where the fuel pressure was increased to raise the equivalence ratio above 0.3. This was done to
increase the pressure in the combustion chamber, sufficiently to separate the boundary layer. It was
observed in shock tunnel experiments undertaken at The University of Queensland’s T4 facility, that
at an equivalence ratio of approximately 0.36, the boundary layer separated. Hence, the flight
experiment provided additional information on boundary layer separation induced by combustion.

6
1.2 Design Overview
The payload assembly schematic is shown in Fig. 1.2.1 and the experimental layout of the supersonic
combustion experiment is displayed in Fig. 1.2.2. A three-dimensional rendered drawing is displayed
in Fig.1.2.5. A picture of the flight module is shown in Fig 1.2.3. The “scramjet” (more precisely
termed a combustion experiment, since the engine creates no thrust) consists of an intake and two
combustion chambers. The thrust surfaces were removed for simplicity. The combustion chambers are
parallel and the cross-sectional dimensions were 9.8x75mm. The intake is a simple wedge of 18
degrees half angle. The intake is 100mm wide. Between the intake and the combustion chamber is a
floor and two side bleeds (see Fig 1.2.4). The bleeds are used to spill the boundary and entropy layers
from the floor and the sidewalls of the intake, as well as allow the shock generated by the leading edge
of the cowl to pass outside the combustion chamber. Thus, an as near as possible to uniform flow
enters the combustion chambers. Therefore, the analysis of the flow path is simplified. Fuel was
injected into only one combustion chamber. This was done to allow fuel-on measurements to be
compared with fuel-off measurements. The fuel injector was a series of holes in the wall of the
combustor (See Fig 1.3.1.). Gaseous hydrogen was the fuel and was injected supersonically through
the holes. The model was designed so that uniform flow would enter the combustion chambers for a
variation in Mach number between 7.2 and 8, as well as at angles of attack between plus and minus 5
degrees.

Figure 1.2.1: HyShot payload schematic

Figure 1.2.2: Experimental layout

Different designs were considered, but it was understood to make the experiment as simple as
possible, and to avoid the difficulties with the large variations in the angle of attack, a significant
amount of spillage from the intake was required. In addition, a simple, but inefficient single stage

7
wedge was used as the intake. It is important to understand, that obtaining benchmark data on
supersonic combustion was one of the main aims of the experiment. Hence, shocks and expansions
entering the combustion chamber would only add unwanted complications. Therefore, by design, they
were removed for all expected flight conditions.

Close tolerance on the dimensions in the region of the intake to the combustor also had to be
maintained to ensure that the shocks and expansions from the cowl and inlet did not enter the
combustor. During re-entry, if unchecked, aerodynamic heating causes the structure to expand and
these tolerances would not be maintained. Hence, the intake and combustor were pinned in this region
to minimize this effect. These pins are shown in Fig. 1.2.4.

Figure 1.2.3: HyShot payload.

Figure 1.2.4: Inlet and combustion chamber entrance.

8
In addition, the unsteady aerodynamic heating which was to be experienced, caused some difficulties
in the design. Although the copper structure of the scramjet helped to dissipate the heat (no active
cooling was used), a gradient in temperature is setup across the plates forming the combustor’s walls.
Typically, across the 12mm plate of the combustor walls, there would be a difference of 80 degrees
Celsius between the inner and outer surfaces, towards the end of the flight. Unless the structure was
suitably constrained this would bend the plate, as it would act like a bimetal strip. Hence, particular
attention was made at constraining the walls of the combustor and ensuring the design was symmetric.
In addition, the model was expected to expand approximately 3mm in its length by the end of the
flight. Expansion joints and slotted holes were incorporated so as to allow the model to move in a
defined manner.

At the front of the scramjet, Pitot probes and intake pressure gauges are mounted (not shown on Fig.
1.2.5). In addition, a cold gas thruster, used to eject the nose cone and re-orient the payload is
installed. A magnetometer is also mounted onto the front section of the scramjet (see Fig. 1.3.2).

Pressure measurements were mainly down the centreline of both combustion chambers. The
transducers were mounted as close as possible to the surface of the combustor to reduce the response
time. However, they were sufficiently far away so that their operational temperature during the
experiment was not exceeded.

The can was the cylindrical section downstream of the scramjet. It contained the gaseous hydrogen
fuel as well as gaseous nitrogen, which was used for the cold gas thruster. The nitrogen was contained
in two, three litre containers and the hydrogen was contained in one, three litre container. The can also
contained the batteries, onboard computer, additional power supplies, relays, telemetry system and
instrumentation such as accelerometers, horizon sensors, pressure and temperature gauges on the
nitrogen and hydrogen supplies. In addition, the can supported the lug, which in turn supported the
payload and the second stage motor whilst on the launcher rail. Installed in the can was also an
umbilical hatch, which was used as the access port for refuelling, communication and power whilst on
the ground. The length of the scramjet and can is approximately 1.4m.

Figure 1.2.5: Three-dimensional view of the HyShot I and II payload.

9
1.3 Instrumentation Details
Instrumentation carried by the payload included pressure transducers, ceramic thermisters (thermo-
resistors), thermocouples, accelerometers, magnetometers and horizon sensors. The onboard battery
voltage was also measured. It should also be noted that all pressure measurements made in the
scramjet were absolute.

Figure 1.3.1: Instrumentation layout and injector location in the fuelled combustion chamber (all
measures in millimetres)

Fourteen pressure measurements were made in each combustion chamber. Fig 1.3.1 displays the
positions of the transducers relative to the leading edge of the combustion chamber. The transducers
were mounted on the inside wall of the combustion chamber. It is important to note that the
combustion chamber outside wall extends 58.7mm upstream of the remaining 3 walls of the
combustor to form the cowl. The combustion chamber leading edge is defined to be in the plane of the
leading edges of the three walls and not at the leading edge of the cowl. Thirteen of the pressure
measurements were made down the centreline of the inner wall of each combustion chamber. The first
measurement was made 101.6mm from the leading edge of the combustor, and the 12 subsequent
measurements were made at 22mm intervals. An additional measurement was made 25mm off the
centreline, 290.6mm from the leading edge of the combustor (see Fig 1.3.1). The fourteen pressure
measurements on the fuelled combustion chamber were designated PA1-PA14, respectively. PA1-
PA13 were located sequentially down the centreline as shown in Fig 1.3.1 (the prefix PA has been
dropped in this Figure). PA14 is located to the side as an aid to the detection of three-dimensional
effects Similarly the unfuelled combustion chamber has pressure transducers PB1-PB14. As can be
seen from Fig. 1.3.1, PA14 and PB14 are on opposite sides of their respective plates. However, when
the plates are fitted to the model, as the plates are inverted with respect to one another, these
transducers are on the same side of the scramjet. To increase the accuracy of these measurements,
different transducer sensitivities were chosen for different locations within the two combustion
chambers. In general, the sensitivity decreased with distance down the combustion chamber, and
furthermore, for the same location in the combustion chamber, the sensitivity was higher in the
unfuelled chamber. All transducers measured absolute values and were temperature compensated.

The pressure transducers were mounted in a stainless steel stem which was screwed and sealed to the
combustion chamber inner wall. The close location of the transducer to the wall of the combustion
camber was chosen to provide a fast response time. This was necessary because during the experiment,
the spin rate was between 5Hz and 7Hz and with angles of attack between 3o and 6o, swings in the
pressure of the order of 30% were expected every 70ms. However, it was also sufficiently far enough
from the wall so as to provide thermal protection to the transducer during the duration of the
experiment.

10
One temperature measurement was made on the cold side of each of the inner combustion chamber
walls using a thermo-resistive sensor. They were located 279.6mm from the combustor leading edge
and 20mm off the centreline and were on the opposite side to the 14th pressure transducer. (See Fig.
1.3.1). The temperature measurements made on the fuelled and unfuelled combustion chambers were
designated as ETA and ETB, respectively.

Figure 1.3.2: Leading edge with Pitot probe and magnetometer location (all measures in
millimetres).

Two Pitot probes were mounted on either side of the intake. Each had two transducers attached to
them (see Fig. 1.3.2 for details of location). These transducers had different ranges so as to cope with
the variations in atmospheric pressure during re-entry. These transducers were designated PPA1,
PPA2, PPB1 and PPB2. PPA1 and PPB1 were the high sensitivity gauges and were appropriate for
pressures up to 101kPa. PPA2 and PPB2 were used from 80kPa and were linear to 350kPa. PPB1 and
PPB2 were located on the same side as the thruster nozzle.

In addition, pressure measurements were also made on the intake. These measurements were made at a
wetted distance of 59.4mm from the leading edge of the intake and 3.6mm off the centreline. These
transducers were designated IPA and IPB. IPA was on the intake to the fuelled combustion chamber,
while IPB was on the intake to the unfuelled combustion chamber.

To control and monitor the fuel flow rate, a pressure transducer was mounted between the fuel control
valve and the injectors. It had a range of 0-1.4Mpa and was designated as flp (fuel line pressure). It
was temperature compensated.

The nitrogen bottle and the hydrogen bottle pressures were also monitored (these gauges were not
temperature compensated). In addition, the hydrogen temperature was recorded using a thermocouple.
The hydrogen bottle pressure and temperature were designated FTP and FTT, respectively. The
nitrogen bottle pressure was designated N2P. All pressure transducers were manufactured by Sensym.

11
1.4 Fuel system
The fuel supply was gaseous hydrogen. It was contained in a single three litre container. A needle
valve operated by a stepper motor controlled the flow of fuel to the injectors. The absence of a
regulator in the system meant that the fuel control system had to cope with low fuel flow rates when
the tank pressure was high (approximately 21MPa), and high fuel flow rates at the end of the
experiment when the majority of the fuel had been depleted and the pressure in the fuel tank was low
(approximately 10MPa). The fuel line pressure was increased in response to the rise in freestream
pressure (and therefore maintaining a constant equivalence ratio).

While the flow in the combustion chamber was supersonic, the injectors remained choked. The fuel
injectors were mounted in the combustion chamber’s inside wall (i.e. the wall along which the
pressure transducers were mounted). Fuel flowed through four holes equally spaced spanwise across
the combustion chamber (see Fig 1.3.1). Each hole had a diameter of 2mm and with its axis
perpendicular to the flow. The flow rate through the injectors can be determined by means of a
pressure transducer located between the needle valve and the injectors (flp) or by the pressure and
temperature measurements made in the hydrogen tank. These tank measurements were made at the
head of the fuel tank.

A calibration constant was determined which related the fuel line pressure (flp) with the fuel mass
flow rate. Experiments were performed where the fuel line pressure was held constant (by controlling
the opening of the valve), while the fuel tank pressure and fuel tank temperature fell. The fuel flow
rate was shown to be proportional to the fuel line pressure between the valve and the injectors.
Although the temperature of the fuel in the fuel tank was falling, measurements of the fuel line
temperature downstream of the valve revealed that the temperature of the fuel remained within a few
percent of the surrounding plumbing temperature. So essentially the valve and associated plumbing
would reheat the fuel as it flowed through them. Furthermore, in the experiments performed on the
bench, this temperature was approximately constant during the short period (7 seconds) that the valve
was opened. This was presumably because the plumbing had a large thermal mass and high thermal
conductivity. Notwithstanding, because the fuel was being delivered to the injectors at a constant
temperature, the mass flow rate through the injectors was essentially proportional to the fuel line
pressure. Furthermore, the temperature dependence, and therefore flow rate dependence, which would
have resulted from the decreasing total temperature of the fuel had the system be insulated, was not
apparent.

The mass of fuel in the fuel tanks was determined from the ideal gas relationship and the pressure and
temperature measured at the head of the fuel tank. The volume of the fuel tank and plumbing up to the
fuel valve was measured at 3.08 litres. It was assumed that these pressures and temperatures are their
total values. This assumption is consistent with the measurements made during calibration. The fuel
line pressure was controlled and at a constant level throughout the calibration process. It had already
been observed that the temperature in the plenum chamber just prior to injection was essentially
constant, thus under these circumstances the flow rate from the fuel tank should be constant.

Six experiments were performed with fuel line pressures between 400kPa and 800kPa to determine the
correlation between mass flow rate and the fuel line pressure. These results were averaged. If it is
assumed that

Mass flow rate of H2 = K * flp,


then
K=5.291e-9 ± 4% Kg/s/Pa.

12
2 HEG Facility and Ground Test Model

2.1 The High Enthalpy Shock Tunnel Göttingen (HEG)

Figure 2.1.1: Schematic of the High Enthalpy ShockTunnel Göttingen (not to scale).

The free piston-driven shock tunnel HEG is operated in reflected mode. It is designed to provide a
pulse of gas to a nozzle at stagnation pressures of up to 200 MPa, and stagnation enthalpies of up to 24
MJ/kg (see e.g. Eitelberg (1994), Hannemann et al. (2000)). Since its commissioning in the early
1990s, HEG has been utilized in numerous space programs. The research activities which have always
been strongly linked with CFD investigations range from the calibration process of the facility and the
study of basic aerodynamic configurations, which are well suited to investigate fundamental aspects of
high enthalpy flows, to the investigation of complex re-entry configurations (Hannemann (2003)).

A schematic of the facility is shown in Figure 2.1.1.The overall length of HEG is 60m and it consists
of three main sections: the driver, consisting of an air buffer and a compression (driver) tube, the
shock or driven tube and the subsequent nozzle/test section. The compression tube is separated from
the adjoining shock tube by a stainless steel main diaphragm. During a run, HEG uses the high
pressure air buffer to drive a free piston down the 33m long compression tube compressing quasi
adiabatically a driver gas mixture of helium and argon in a transient process, allowing generation of
high driver temperatures and pressures. The driver pressure is regulated by the bursting of the scored
main diaphragm. After diaphragm burst, a strong shock wave propagates down the 17m long shock
tube and is reflected from the end wall, creating a region of high pressure, shock-heated test gas. When
this nozzle reservoir region is formed, the secondary diaphragm, a thin mylar sheet, ruptures and the
test gas expands through the convergent-divergent hypersonic nozzle to provide the free stream flow.

The extension of the HEG operating range to allow the ground based testing of complete scramjet
engines included the design and construction of a contoured nozzle (HEG nozzle III) which provides
parallel Mach 7.8 flow with a test core of 400mm diameter, the construction of a new 851kg piston as
well as the installation of a fuel system for the delivery of hydrogen fuel to the wind tunnel model.
Finally, new test conditions at total enthalpies of about 3MJ/kg were designed (Gardner et al. (2004)).
These nominal conditions are summarized in Table 1.

Two operating conditions have been used for the present HyShot postflight analysis. These conditions
correspond to nominal flight altitudes of 32.5km and 27.1km. The corresponding conditions in the
Hyshot flight data were chosen by matching the Pitot pressure during the ground test with that in
flight, and then setting the angle of attack and fuel mass flow rate required to their values at that Pitot
pressure during the flight. These HEG operating conditions are summarised in Table 2.1.1.

13
Nominal Nozzle Stagnation Nozzle exit Nozzle exit Nozzle exit Nozzle Nozzle Nozzle
flight stagnation Enthalpy static temperature Pitot exit exit exit unit
altitude pressure [MJ/kg] pressure [K] pressure Mach density Reynolds
[km] [MPa] [Pa] [kPa] number [kg/m3] number
32.5 6.90 3.51 709 266 62 7.8 0.0093 1.4 106
27.1 16.68 3.20 1711 242 146 7.8 0.0246 3.8 106

Table 2.1.1: HEG operating condition XI (27.1 km) and XII (32.5 km) used for the HyShot postflight
analysis.

Computational Fluid Dynamics (CFD) was performed with the DLR CEVCATS-N code (Brück et al,
1997), for conditions XI (27.1km) and condition XII (32.5km). The comparison of measured and
computed Pitot pressure and stagnation point heat flux profiles at the nozzle exit are shown in Figure
2.1.2. The exit Mach number for both conditions is 7.8 (Figure 2.1.3) and it varies less than 2%
between both conditions (Figure 2.1.4).

Figure 2.1.2: Comparison of computed and measured Pitot pressure and stagnation point heat flux
profiles for HEG conditions XI and XII (the measured data is averaged from 6ms to 9ms after shock
reflection).

14
Figure 2.1.3: Mach number contour diagram (CFD) for the HEG nozzle III at condition XI.

Figure 2.1.4: Contours of Mach number fractional difference between CFD for conditions XII and XI.

The fuel system on HEG consists of a 12mm diameter Ludwieg tube (Figures 2.1.5 and 2.1.6), and a
fast acting solenoid valve (not shown in Figure 2.1.6) positioned within the model. The Ludwieg tube
can deliver a pulse of fuel with constant pressure for 13-20ms.

The fuel is injected into the combustion chamber of the fuelled side of the model in the same way as
for the flight model. A pressure transducer monitors the pressure in the fuel line, and the pressure in
the Ludwieg tube is monitored before and after the test.

The injectors are calibrated by filling the Ludwig tube with high-pressure fuel, and then opening the
fast acting solenoid valve, allowing fuel to escape through the injectors. A calibration constant is
found,

 V 
K =  L 
∆PL
 RTL 
γ −1 γ +1
,
2γ 2γ
P L P
fl ∆t

where VL, TL, and R are the volume (2.00L), initial temperature (300K) and gas constant (4128.1
J/kg/K) respectively in the Ludwig tube. The valve opens, and then closes within a short time (∆t),
typically between 10ms and 50ms. ∆PL is the change in pressure in the Ludwieg tube over that time.
PL, measured by gauges 7 and 8 in Figure 2.1.6, is the initial pressure in the Ludwieg tube, and Pfl is
the average pressure in the reservoir behind the injectors. The instantaneous pressure (Pfl) is recorded
with pressure transducer FLP within the model (not shown in Figure 2.1.6), and the average is found
by,

15
∫P
γ +1
γ +1 2γ
2γ fl dt
P = .
∆t
fl

γ +1 γ −1
The powers 2γ
and 2γ
are corrections (from a power 1) for an isentropic drop in pressure and
temperature of the fuel in the Ludwig tube over time (∆t), due to its finite size.

The calibration constant K is used for each test to find the instantaneous mass flow rate of fuel ( m
D hy )
for a given Ludwieg tube filling pressure and static pressure measured in the reservoir behind the fuel
injection ports,
γ −1 γ +1
2γ 2γ
mC hy = KP L P fl .

The equivalence ratio is,


8mD hy
φ= ,
mD ox

where m  ox is the mass flow rate of oxygen into the engine. m ox is calculated by the engine geometry
and the freestream parameters. The fuel system was calibrated for a range of pressures. The calibration
constant
K= 4.85e-9 ± 3% kg/s/Pa
was determined. All of the data for the fuel calibration is included with this report.

Figure 2.1.5: Photograph of the Ludwieg tube installation on HEG used for the H2 injection system.

16
Figure 2.1.6: Schematic of the H2 injection system installed on HEG .

17
2.2 Hyshot Ground Test Model Design
The ground test model represents the 1:1 scale geometry of the flight model. Due to the low integral
heat loads in the HEG the use of special materials to build the model was not required and a free
machining steel (ST37K) was used. The number of pieces was reduced, as no separate tips were
required for the leading edges, and holes previously slotted to allow expansion were made round. The
two chambers, joined together in the flight model, were made separable so that they could be given
equal, positive angles of attack during the tests (Figure 2.2.1). In order to reduce the manufacturing
time, only the inside surfaces were milled to a good finish.

The model mounted in the HEG test section is shown in Figure 2.2.2. Here, the static pressure probe,
the fuel delivery pipe, fuel line pressure gauge mount, and the shielding for the transducer cables are
also visible.

Figure 2.2.1: Side assembly of the Hyshot wind tunnel model.

18
Figure 2.2.2: Hyshot wind tunnel model mounted in the HEG test section. Both chambers are at an
angle of attack of +3.6°.

19
2.3 Instrumentation of Ground Test Model
The instrumentation of the ground-test model (shown in Figure 2.3.1) was designed to be a superset of
the instrumentation of the flight model.

In the combustion chamber (fuelled and unfuelled) 15 pressure transducers are used, with an additional
two pressure transducers on the intake. Of the pressure transducers in the combustion chamber 13 are
on the centreline at the same positions as those in the flight model (Figure 1.3.1), with one extra
transducer at the same spacing continuing further downstream on the centreline, and one off centre in
the same position as in the flight model. The transducers along the centreline are named FP01 to FP14
for the fuelled chamber and UP01 to UP14 for the unfuelled chamber. The off-centre transducers are
called FP15 and UP15 for the fuelled and unfuelled chambers respectively. There are two pressure
transducers in the inlet for each engine, 79.8 and 119.8 mm from the leading edge (measured along the
wetted length). These are called FP16 and FP17 in the fuelled engine inlet, and UP16 and UP17 in the
unfuelled engine inlet. In addition, the fuel line pressure is monitored (FLP), as is the Pitot (pt2) and
static (static) pressure in the freestream. The ranges of the pressure transducers in the Hyshot wind
tunnel model are given in Table 2.3.1.

On the model, 27 thermocouples are mounted. These are installed, along a line 6mm off the centreline,
equidistant from the two proximate pressure transducer positions. Not all of the positions shown in
Figure 2.3.1 were used. The first 10 centreline positions (FH04 to FH13) were used in the fuel-on
chamber, and the first 7 (UH04-UH10) in the fuel off chamber. There were three heat transfer gauges
(FH01-FH03 and UH01-UH03) placed in each combustion chamber 28.13 mm from the centreline,
directly behind one of the outer injector ports, at 11.5mm, 30mm and 50mm behind the injector port.
There are two thermoelements in the inlet for each engine, 99.8 and 139.8 mm wetted distance from
the leading edge. These are called FH18 and FH19 in the fuelled engine inlet, and UH18 and UH19 in
the unfuelled engine inlet.

Figure 2.3.1: Schematic of the positions of the sensors in the Hyshot wind tunnel model.

Sensors Type Range*


FP01-FP12, FP15, FLP Kulite XCEL-100 0-7 bar
FP13,FP14,FP16, UP01-UP03,UP16 Kulite XCEL-100 0-3.5 bar
UP04-UP15, UP17, FP17 Kulite XCEL-100 0-1.7 bar
*Factory calibrated range is as above

Table 2.3.1: Pressure sensor ranges of the Hyshot wind tunnel model used in the HEG.

20
3 HEG Ground testing
3.1 Methods for the analysis of results
In order to show the method of analysis of an experiment in HEG, the example of test 666 will be
used. In this experiment, HEG operating condition XII (corresponding to 32.5 km flight altitude, see
Table 2.1.1) was used. The angle of attack α is 3.82°, meaning that the inlet is at an angle of
3.82°+18°=21.82°. Further the data resulting from run 663 using HEG operating condition XI
(corresponding to 27.1 km flight altitude, see Table 2.1.1) will be discussed in this section.

The first step of the data analysis is to select the usable test time. The period of constant HEG flow
conditions in the test section is selected by comparing the normalised values of the nozzle stagnation
pressure with the Pitot pressure, and static pressure taken from the permanent probe measurements in
the test section. In Figure 3.1.1 the nozzle stagnation pressure (STP#2_norm) is time-offset with
respect to the Pitot pressure (pt2_norm) and the static pressure (pstat_norm) to account for the time
delay for the air to pass through the nozzle. This offset is on the order of one millisecond for the Mach
8 nozzle used for the present tests (HEG nozzle III). The time is set to be zero when the primary shock
in the shock tube is reflected from the endwall.

Between 4.5ms and 8.0ms the Pitot pressure follows the stagnation pressure closely. This is an
indication that the expansion ratio in the nozzle is a constant, and as the expansion ratio is set by the
thickness of the boundary layer, the differences between these traces before 4.5ms indicates that the
boundary layer in the nozzle is not yet fully established. Thus the test time starts after 4.5ms for test
666. At 8.0ms the static pressure begins to drop away from the other two traces indicating that the test
gas is becoming contaminated with the argon/helium driver gas. Consequently, an established flow in
the HEG test section was obtained between 4.5-8.0ms for run 666. In order to determine the test time
used to evaluate the data measured in the combustors it must be assured that the fuel pressure is
constant and that the fuel flow is fully established. Due to the fact that in the case of test 666 (Figure
3.1.2), the fuel pressure started to drop at 7.5ms, only the period 4.5 to 7.5 ms can be used as test time.
In addition to these criterions, the selection of the test time depends on the properties of the scramjet
engine. Typically, the flow in the model requires two flow lengths for attached turbulent boundary
layers and three flow lengths for attached laminar boundary layers to be established (Anderson, 1990).
If separated regions are present in the flow, the required flow lengths for the flow establishment would
be increased to about 20-50. However, not all of this required time must be during the test time. The
data obtained in the HyShot model shows that the flow establishment and the establishment of the
combustion typically requires no longer than 1 ms. The pressures in the combustors are stable by the
time that the nozzle boundary layer is established (see for example Figure 3.1.3, transducer UP09).
The trace of transducer FP09 in Figure 3.1.3 also indicates that the combustion process becomes
unstable when the Pitot pressure is reduced at the end of the test time (after 10ms after shock reflection
in Figure 3.1.3). The reason for this is that with decreasing Pitot pressure the equivalence ratio
increases and consequently the pressure increases. This additional pressure increase can result in an
unsteady flow in the combustor due to boundary layer separation.

The operation of a piston driven shock tube is different to that of a continuous facility in that a desired
flow condition cannot always be reproduced exactly in the test section. Due to the pulse nature of the
facility no tuning of the condition during a test run is possible. The determination of what test
conditions were achieved is determined from the test data after the experiment is completed.

In order to determine the flow conditions for one particular run, the conditions in the nozzle reservoir
are first determined by using the shock speed (SS in Table 3.1.1) and filling pressure in the shock tube
(Pfill in Table 3.1.1) as input for a shock tube calculation assuming that air is in equilibrium. This
calculation is performed using the program ESTC (McIntosh, 1968). The shock speed is measured in
HEG by the combination of five co-axial ionisation gauges in the shock tube spaced at approximately
3.3-meter intervals along the length of the shock tube, and the stagnation pressure gauges to get the
correct shock speed at the nozzle end of the shock tube. ESTC calculates an ideal value for the

21
pressure and temperature achieved behind the reflected shock assuming a one-dimensional shock. This
gas is adiabatically expanded to match the measured nozzle stagnation pressure (Pres in Table 3.1.1),
which compensates for the three-dimensional nature of the flow.

HEG 666, normalized pressures


1.50

STP#2_norm
1.25
pt2_norm
pstat_norm

1.00
pressure [ ]

0.75

0.50

0.25

0.00
0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0 16.0

time [ms]

Figure 3.1.1: Tunnel standard pressures for the selection of the test time in HEG, run 666.

HEG 666 fuel timing data


Magnitude

Pt2
FLP

0 4 8 12 16

Time [ms]

Figure 3.1.2: Normalised fuel and Pitot pressures for the selection of the test time for run 666.

22
Magnitude

Pt2
UP9
FP9

0 5 10 15

Time [ms]

Figure 3.1.3: Normalised pressures in the combustors of the HyShot model and free stream Pitot
pressure in HEG.

For run 666 the shock speed was measured to be 1839 m/s, where the shock tube filling pressure was
43.3kPa. The calculated pressure and temperature behind the reflected shock are then 3104K, and
10.5MPa. This gas is then expanded to 2880K and 6.895MPa.

The remainder of the calculations of the freestream data are made using the stagnation enthalpy. The
temperature in the freestream (Tf in Table 3.1.1) can then be calculated from the definition of the Mach
number, and the total enthalpy to be:
H0 / Cp
Tf = ,
M 2γR
1+
2C p

where Cp=1004, R=287 and γ=1.4 for air. For shot 666 this temperature is 266K. The velocity (vf) can
then be calculated from the definition of the Mach number, to be 2548m/s.

As can be seen in Figure 2.1.2, if the Pitot pressure is taken from a point 190mm from the axis of the
nozzle then the measured Pitot pressure will overestimate the mean Pitot pressure. In order to use the
Pitot pressure to make other calculations, an effective Pitot pressure is calculated by multiplying the
measured values by KPt2=0.9228 and KPt2=0.9065 for condition XI and XII, respectively. If the Pitot
pressure is known, and the velocity has been calculated as above, then the density (ρ in Table 3.1.1)
follows from the Rayleigh Pitot relationship:
 
  γ +1  γ −1 
γ

 1  ⋅M2 
= 2  2 
1 
Pt 2 ⋅ K Pt 2 ⋅ ⋅ ρv 2 .
 γM  2γ γ − 1  γ −1 
   
  
M2 −
 γ + 1 γ + 1
For run 666 the resulting value for the free stream density is 9.3g/m3. Using the calculated free stream
temperature and density, a static pressure of 709 Pa results for run 666. Due to currently not resolved
uncertainties concerning the calibration of the static pressure probe in HEG at the new low enthalpy
operating conditions, the static pressure is taken to be that calculated through the Pitot pressure, rather
than the one measured when static pressure is required as an input for the data evaluation of the flow

23
in the HyShot combustors. Therefore, the measured static pressure is only used in a qualitative way
here in the framework of the test time selection process.

A relation exists between the stagnation point heat transfer, and the total enthalpy. This may be used
as a check on the enthalpy calculated through the shock speed above. The correlation of Verant:

 H − H wall 
⋅ 23.787 ⋅  0 
1.0688
Pt 2eff
 R ⋅T 
qD =
 
,
rSphere ref

where rSphere=0.01, Hwall=CpTwall=1004*300, R=287, and Tref= 273.15K, has been shown to be suitable
in HEG at higher enthalpies (the comparison between the values calculated through this method agree
with the reservoir enthalpies determined by ESTC computations). The nozzle reservoir enthalpy for
test 666 resulting from ESTC was 3.51 MJ/kg and the calculated freestream enthalpy using the Verant
correlation (H0-qdot in Table 3.1.1) was 3.44MJ/kg, a difference of about 2%.

The results of the calculations of the conditions in the freestream of the HEG are summarised for all
runs performed for the present investigations in Table 3.1.1.

Run Pfill (kPa) SS (m/s) Pres (MPa) H0 (MJ/kg) Mf Tf (K) vf (m/s)


662 140 1720 16.49 3.00 7.8 227 2356
663 140 1804 16.68 3.20 7.8 242 2432
664 140 1770 16.20 3.10 7.8 235 2395
665 43.3 1885 7.30 3.68 7.8 278 2606
666 43.3 1839 6.90 3.51 7.8 266 2548
667 140 1743 16.70 3.07 7.8 232 2380
668 43.3 1807 6.79 3.42 7.8 259 2515
669 140 1755 16.72 3.09 7.8 234 2391
670 140 1747 17.08 3.09 7.8 234 2390
671 140 1759 16.79 3.10 7.8 235 2395
672 140 1758 17.06 3.11 7.8 236 2399

Run Pt2 (kPa) ρ (kg/m )


3
Pstat-calc (Pa) qC (MW/m2) H0-qdot (MJ/kg)
662 144.1 0.0259 1687 3.83 3.01
663 146.1 0.0246 1711 3.66 2.88
664 138.8 0.0241 1625 3.61 2.91
665 64.4 0.0093 741 3.11 3.59
666 61.7 0.0093 709 2.91 3.44
667 143.5 0.0253 1680 3.71 2.94
668 60.3 0.0093 694 2.87 3.44
669 143.1 0.0250 1676 3.76 2.98
670 145.5 0.0254 1704 3.70 2.91
671 145.0 0.0252 1698 3.70 2.92
672 145.6 0.0252 1706 3.90 3.05
Table 3.1.1: Freestream calculations for the runs in HEG performed during the present investigations.

The trajectory points of the HyShot flight selected for comparison with the ground test data were
chosen to have equivalent Pitot pressures with condition XI and condition XII. In flight the angle of
attack is changing quickly with time and thus the angle of attack for comparison with the tunnel data is
the instantaneous angle of attack at the reference Pitot pressure. Due to some variation between tests at
the same condition all results were normalised with the Pitot pressure. Eleven runs in HEG were
performed for this contract. The runs are summarised in Table 3.1.1. Each run used two test chambers
of the HyShot scramjet geometry which were independently assigned angles of attack (α). The side
with fuel injectors in the combustion chamber is denoted side A, and the side without injectors is

24
denoted side B. There were eight tests at condition XI and three tests at condition XII. The conditions
for side A of the scramjet model are summarised in Table 3.1.2.

Table 3.1.2: Calculated conditions through the HyShot configuration for each condition tested.
Pressures normalised to Pt2 are in grey in the row below the absolute pressures.

A typical pressure distribution looks like Figure 3.1.4. Chamber A has fuel injected into it, and
chamber B has no fuel. The difference in the pressure traces due to combustion is clear. The points to
the left of the break in the x axis are on the intake of the engine. The pressures on the inlet and in the
combustion chamber may be predicted analytically by using the oblique shock relations. For run 666
the flow on the inlet has been turned through an angle of 21.82°. The flow in the combustion chamber
has been turned through an additional 18°. The shock structure of the HyShot wind tunnel inlet is
shown in Figure 3.1.5 and the analytical calculations of flow data through the two oblique shocks are
included in Table 3.1.2 for all tests undertaken and for the flight conditions corresponding to
conditions XI and XII.

25
Figure 3.1.4: Normalised engine pressure distribution, test 666.

Figure 3.1.5: Shock structure of the HyShot wind tunnel model inlet. 1=Freestream, 2=Inlet,
3=combustion chamber.

A steady rise in the pressure in the combustion chamber is observed due to the thickening of the
boundary layer causing a constriction of the supersonic flow.

The comparison of the data in Figure 3.1.4 with the calculations in Table 3.1.2 yields agreement
between the analytically calculated pressures in the inlet and in the front of the combustion chamber,
using the assumption that the pressures nearest the front of the combustion chamber most closely
reflect the inviscid prediction of pressure. The pressures in the fuel-on trace show a slight increase
which is associated with the injection of the fuel into the duct.

26
The fuel in the HyShot scramjet is injected from four portholes equally spaced across the duct. As the
fuel is injected, the formation of a curved shock around the front of the stream can be expected. The
fuel is turned downstream by the main flow. The penetration of the fuel into the main flow can be
estimated by the Newtonian method of Schetz & Billig (1966), shown in Figure 3.1.6. The penetration
will be the same for both of the conditions tested as it is a function of the ratio of dynamic pressures of
the freestream and fuel, which is the same for both conditions investigated. In Figure 3.1.6, the
calculated penetration distance of the fuel against downstream movement for an equivalence ratio of
φ=0.30 is plotted. It is obvious that the fuel penetrates approximately half way across the combustor
duct which has a height of 9.8 mm. This clearly indicates that sufficient fuel enters the freestream that
it can be expected that the majority of the combustion process takes place within the supersonic region
of the duct.

Figure 3.1.6: Penetration depth of hydrogen fuel into the freestream.

The injection of fuel slows the freestream flow, and cools it down. This effect is calculated in Table
3.1.2. The temperature after the cooling due to the fuel is used as the starting value for the calculations
of the initiation and combustion times of the fuel.

A departure between the values measured in the fuel-on chamber and the fuel off chamber occurs
between transducers FP03 and FP04. This rise is due to heat release from combustion. The initiation
phase of the combustion is characterised by endothermic chemical reactions which do not correspond
to any large pressure increase. The reaction after that is characterised by a sharp pressure rise. The
initiation time (τi) and the combustion time (τr) are predicted by Huber et al (1979) as:

8 × 10 −9 e 9600 / T
τi =
P × 10 5
1.05 × 10 −4 e −1.12T / 1000
τr = .
(P × 10 ) 5 1.7

It is emphasised here, that these correlations are strictly true only in a premixed flow. Using the local
flow velocity and converting the initiation time into a length, it is bounded in the present experimental
results by the fuel injection location and the location of the first pressure rise. The predicted rise time
for shot 666 is 15.96µs, resulting in an initiation length of 31mm. This would mean that the pressure
rise due to combustion should begin upstream of the location of the first pressure transducer. The
difference with the experimentally observed initiation length must be due to the time taken for the
mixing to occur. Due to the injection of cold fuel in the present experiment, combustion will be
inhibited in fuel-rich regions of the flow. From Figure 3.1.4 it can be seen that a mixing length of 58-
80mm is required in run 666. The combustion time predicted for test 666 is 73.7µs, or a combustion
length of 140mm.

27
The cooled flow after fuel injection is also used as the starting value for an equilibrium combustion
calculation. This gives the maximum pressure in the combustion chamber if all of the fuel combusts,
ignoring the compression due to the construction of the boundary layer. Previous experiments have
shown that in the case where all of the fuel can be expected to react during the residence time in the
combustion chamber, that this equilibrium pressure predicts the experimentally found pressure within
5% if no boundary layer separation occurs (Gardner et al. (2002)). However, it should be noted that
these results are valid for combustion chambers where the boundary layer is comparatively thinner to
the present ones. The predicted pressure for equilibrium combustion for HEG run 666 (Table 3.1.2) is
2.6 times the Pitot pressure of the freestream.

Separation of the boundary layer by an adverse pressure gradient in the combustion chamber, resulting
either from a strong shock or combustion has been found to be well predicted (Paull (1999)) by the
relation of Korkegi (1975):
P
= 1 + 0.3M 2 M ≤ 4 .5
Pi
and
P
= 0.17 M 2.5 M ≥ 4 .5
Pi
Korkegi’s separation pressure is noted for test 666 in Table 3.1.2, and is 2.39 times the Pitot pressure.

Separation could lead to a strong constriction of the flow resulting in engine unstart. That means that a
normal shock is created, which subsequently could move upstream choking the combustion chamber.
The pressure behind a normal shock can be determined according to:
 2γ 
P = Pi 1 + (M 2 − 1) .
 γ +1 
In Table 3.1.2 this value is shown for each run. For run 666, a normal shock inside the fuel-on duct
would generate a normalised pressure ratio of 5.7. Such a high pressure ratio was not detected during
the test in HEG. Further, the pressure in the duct is increasing as expected by a supersonic flow which
undergoes a constriction. This strongly indicates that the flow in the combustor remains supersonic
with combustion.

The normalised wall heat flux data measured in HEG using operating condition XII is shown in Figure
3.1.7. This data is normalised with the stagnation point heat transfer (q00) measured on a 20.0mm
diameter sphere, positioned in the freestream of the HEG test section. When fuel is injected in the
combustor, two competing effects concerning the heat flux development are present. The injection of
cold hydrogen into the combustion chamber cools the flow in the combustion chamber, however, the
addition of hydrogen also increases the density of the gas in the combustor. In the fuel-off duct, the
rise of wall heat flux 181mm downstream of the combustor leading edge indicates that the boundary
layer undergoes transition from laminar to turbulent flow. In the fuel-on duct an increased heat transfer
is measured already further upstream indicating that tripping of the boundary layer is caused by the
fuel injection. The lower heat flux measured in the first part of the HyShot fuel-off combustor agrees
with the laminar heat flux predicted using a Blasius profile and a wall temperature of 300 K. In Figure
3.1.7 an estimation of the turbulent heat flux level in the fuel-off chamber is also given using the
theory of van Driest II. It should be noted that the heat flux resulting from this correlation was
evaluated using the free stream parameters at the combustor inlet. The continuously changing free
stream conditions of the flow in the duct were not taken into account resulting in a lower turbulent
heat flux prediction by the van Driest II correlation compared to the measurements in HEG. In the
fuel-on chamber, the position of the rise in the heat flux at 159mm corresponds with the position
indicated by the rise of the pressure due to combustion (see Figure 3.1.4).

28
From a number of previous experimental investigations in scramjets (for example Paull, 1999), it is
expected that transition will take place between Reynolds numbers of 5×105 and 2×106. The results for
the HyShot model obtained in HEG show that transition occurs in the combustion chamber between
Reynolds numbers of 7.6×105 and 1.0×106.

The spike in heat transfer at the front of the fuel-on combustion chamber appears to be from the third
reflection of the shock formed at the tip of the cowl. The position of the rise in the heat flux data is
between FH04 and FH06, agreeing with the position indicated by the pressure data.

Figure 3.1.7: Normalised engine heat flux distribution, HEG run 666.

In Figure 3.1.8 the distributions of the engine pressure, normalised with the Pitot pressure of the free-
stream, obtained from the measurements in HEG (run 666) and in flight are shown. It is clear from
Figure 3.1.8 that the normalised pressure distributions obtained in flight, revealing a significant
pressure rise in the fuel-on combustion chamber compared to the fuel-off chamber, can be
qualitatively reproduced by the ground based tests in HEG. However, a more detailed comparison of
the flight and HEG data shows that quantitative differences in the pressure distributions exist.

29
Figure 3.1.8: Comparison between normalised pressure distributions for HEG run 666 and flight at
32.5 km altitude.

In the fuel-off duct, a continuous rise of the pressure is observed due to the thickening of the boundary
layer causing a constriction of the flow. The points to the left of the break in the x-axis are on the
intake of the engine. Inside the fuel-off duct the pressure level measured in HEG is 20% lower than the
flight pressure level, but the qualitative agreement between both pressure distributions is good. It
should be noted that the application of analytical methods (as discussed above) and CFD confirm the
pressure level measured in HEG (see Figure 3.1.12). This quantitative difference between flight and
ground based testing needs further investigation. A possible explanation is that the geometry of the
cowl and combustion chamber leading edges may have changed during the flight due to ablation.

In HEG fuel was injected into the combustion chamber using an equivalence ratio of 0.45. The
measurements in HEG show that a departure between the values measured in the fuel-on and the fuel-
off chambers occurs 170mm downstream of the combustor leading edge. This rise is due to the heat
released from combustion. In the first part of the ducts, the pressure difference between fuel-on and
fuel-off is higher in flight than in HEG. Further downstream, the pressure increases continuously in
flight and in HEG until a peak value is reached at 280mm downstream of the combustor leading edge.
In this section of the combustors, the flight data still exceeds the values measured in the HEG.
Subsequently the flow expands onto the thrust surface.

The main stream flow conditions after fuel injection in the HEG test were used as the starting value for
an equilibrium combustion calculation. This provides the maximum pressure in the combustion
chamber which would be reached if all of the fuel combusts. A normalised pressure ratio of 2.6 is
predicted for an equivalence ratio of Φ=0.45. The equilibrium combustion pressure exceeds the value

30
measured in HEG by 9%. Due to the fact that the equivalence ratio in flight was 15% smaller than in
HEG, the equilibrium combustion calculation was also performed using the flight equivalence ratio of
Φ=0.38 and the combustion chamber conditions of HEG. Compared to the equivalence ratio Φ=0.45,
the resulting peak pressure value is reduced by 9%. It should be noted that due to the inconsistency in
the pressure in the fuel-off duct obtained in flight, a prediction of the equilibrium combustion pressure
for the flight test could not be made with any certainty.

An adverse pressure gradient in the combustion chamber can cause boundary layer separation. As
discussed above, the critical pressure ratio above which separation might occur can be predicted with
the correlation of Korkegi. This pressure is also given in Figure 3.1.8. It indicates that major flow
separation should not be present in the fuel-on combustor in HEG.

Figure 3.1.9: Normalised engine pressure distributions, HEG run 663.

A typical pressure distribution for condition XI (Figure 3.1.9), shows many of the same features as for
condition XII. In the case of run 663 (with conditions as shown in Table 3.1.2), the rearmost of the
pressure transducers on the inlet showed more noise than normal, probably due to a particle strike.
Apart from this the pressures on the inlet agree with the analytical calculations in Table 3.1.2 as do the
pressures in the start of the combustion chamber. The pressure rise signalling the start of combustion
arrives at the same point as for run 666 as predicted, showing that the mixing lengths for the
conditions are also similar. However a sudden pressure rise is observed where a steady pressure rise
along the length of the duct was expected.

The pressure predicted due to equilibrium combustion is 2.3. As can be seen in Figure 3.1.9, there is a
portion of the flow which exceeds this predicted pressure level, in particular the peak at FP09
(279.6mm downstream the combustor leading edge). This is a strong indication that there is separated
flow in the duct, especially as the predicted Korkegi pressure at which the flow should separate is
2.39.

31
The pressure peak at transducer FP09 appears in many of the experiments in which high pressures are
reached in the combustion chamber. However, a similar peak is not seen in the HyShot data. The
magnitude of this pressure peak increases with increasing equivalence ratio. It should be mentioned
that for the experiment using Φ=0.35, the signal obtained from transducer FP09 went off-scale and the
correct value can not be determined from the measurement. However, the trend of increasing pressure
peak with increasing equivalence ratio can clearly be determined. Further, in the temporal
development of the data, no indication of duct choking is observed. For that reason, even being off-
scale, the data point was not removed from Figure 3.1.9 The origine of this peak is currently not
known. One possible explanation could be that it may be caused by an oblique shock wave generated
at the leading edge of a separation bubble followed by an expansion caused by the decreasing
thickness of the separation bubble and the upstream influence of the expansion resulting from the
diverging duct.

Figure 3.1.10: Normalised engine heat flux distribution, HEG run 663.

The heat flux measurements of HEG run 663 (Figure 3.1.10) show that the boundary layer has started
transition from laminar to turbulent flow upstream of the location of the first thermocouple. If the
boundary layer is transitioning at equivalent Reynolds numbers as in run 666 then the first two
thermocouples should be in the transition region. In fact these two signals show noise and a deviation
from the other transducers in the duct which could be attributed to transition. The heat transfer in the
combustion chamber is three times the laminar prediction and is close to the prediction of van Driest
II. The mixing effect of the cold fuel into the hot freestream is evident in the reduced heat flux
measured by the first 3 transducers (which are positioned directly behind an injection port) in the fuel-
on duct, and by the first transducer on the centreline. As in run 666 the rise in the heat flux data is at
the same position as the pressure rise.

32
Figure 3.1.11 shows the combustion chamber pressures from HEG run 663 compared with the
pressures from the equivalent point in the flight test envelope. Results on the inlet agree between flight
and ground tests. Within combustion chamber B the flight test pressures are 55% higher in the mean
than those measured in the HEG.

Figure 3.1.11: Comparison between normalised pressure distributions measured in HEG run 663 and
the corresponding flight condition at 27.1km altitude.

CFD was performed for run 663 and 666 for the fuel-off combustors (Figure 3.1.12). Computations
were made using the DLR-TAU code and a Spalart-Allmaras turbulence model. It can be seen from
Figure 3.1.12 that the results agree with the experiments obtained in HEG.

Figure 3.1.12: Comparison between normalised pressure distributions for HEG runs 663 and 666 and
computations performed using the DLR-TAU code.

33
3.2 Results and Discussion

Figure 3.2.1: Comparison of results for different equivalence ratios.

Figure 3.2.1 shows a comparison between tests at different fuel equivalence ratios for conditions XI
and XII.

Note that although data from run 665 (condition XII) is included, the experimental data suggests that
the fuel injectors did not function correctly during run 665, and the equivalence ratio may be below the
value quoted in Figure 3.2.1. The pressure rise found in the case of run 665 does correspond with that
calculated in Table 3.1.2. In addition, the boundary layer tripping effect of the injectors can be clearly
seen in the heat flux data for test 665.

For condition XI run 672 is included for comparison as in this run nitrogen was used instead of air as
the test gas. In this way the effect of the injection of hydrogen on the flow can be studied separately to
the combustion effects. The equivalence ratio quoted for run 672 is that equivalence ratio which would
have been achieved if the test gas had been air at the same Pitot pressure. The wall heat flux data
shows a steady increase with increasing equivalence ratio in the main part of the combustion chamber.
In the front of the combustion chamber, higher equivalence ratios correspond to reduced heat flux
measurements as the cooling effect of the hydrogen fuel is increased. The reduced value of the last
heat flux measurement in the combustion chamber corresponds with reduced pressures in this region,
and appears to be due to the influence of the expansion at the end of the combustion chamber.

While the heat flux measurements for condition XI show a positive correlation between heat flux and
equivalence ratio, the pressure distributions for condition XI show the interesting characteristic that

34
three increasing equivalence ratios show three qualitatively different pressure distributions. Run 663
has a rapid rise to a level above the pressure predicted for boundary layer separation followed by a
slow rise. Reducing the equivalence ratio by 10% to that used for run 664, results in a typical profile
expected for supersonic combustion and an agreement in predicted pressure levels with those
presented in Table 3.1.2, with the exception of the peak at FP09. A further 20% reduction in the
equivalence ratio results in the distribution of run 667, where the value of the pressure peak at the
front of the combustion chamber agrees with predictions but drops to be much lower than expected at
the back of the combustion chamber. This behaviour is currently not completely understood and flame
extinction could be one possible explanation. The rise in the heat flux measurements for run 667,
indicates that combustion occurs initially. Further downstream the heat flux drops, however, stays
above the level of run 672. It is likely that the cause of the qualitative differences in the pressure
distributions is the result of the formation of a separation in the combustion chamber at higher
equivalence ratios.

Figure 3.2.2: Comparison of results for different angles of attack.

The pressure trace for chamber A in run 666 where fuel is injected agrees well with the data in flight
(Figure 3.1.8) except at the front of the duct. To remove the possibility that the difference was caused
by a miscalculation of the freestream temperature or angle of attack, it was decided to increase the
temperature in the duct so that the ignition point should moved upstream to between FP01 and FP02.
In order to achieve this the angle of attack was increased to 6.0o. For this condition, the Korkegi
separation pressure and the pressure obtained after a normal shock do not significantly change, but the
initiation time is reduced by a factor of two. Using this method the position of the pressure rise was
successfully moved to the desired position (Figure 3.2.2). As expected, for run 668 at condition XII,

35
the pressure starts to rise immediately downstream the first pressure transducer. However, after the
pressure rises very rapidly, it subsequently drops again. This would generally be an indicator of
subsonic flow in a converging duct. However, other indications support the conclusion that the flow in
the duct is not subsonic: There is a rising pressure in the front of the duct from a starting pressure
which is consistent with the pressure predicted by the analytical method using oblique shock relations,
there is no peak in the pressure distribution which would be expected if there were a standing shock in
the duct and the combustion is stable and not moving forward in the duct. The only possibility that the
flow could show a decreasing pressure and still being supersonic, is if the duct was not parallel. This is
an effect which would be obtained from the formation of a long separation bubble downstream the
initiation of combustion. Since Korkegi’s separation pressure is being exceeded in run 668, the
formation of a separation region is the most likely explanation of the obtained result. Additionally, it is
impossible that the observed pressure could have been achieved solely due to combustion. It is
interesting to mention that the pressure distribution of run 668 looks qualitatively very similar to that
found in the flight experiment at 27.1km altitude (Figure 3.1.11). The heat fluxes in Figure 3.2.2 do
not show any rise until after FH03, which is consistent with the pressure data. There is no evidence
that a large separation region is shielding the gauges from the temperature rise in the freestream.
However if the boundary layer were to separate, it would separate first on the cowl due to the higher
Reynolds number there.

When the angle of attack is increased for condition XI (Figure 3.2.2), no such qualitative change
occurs. The pressure distribution for run 670 shows an earlier rise than that seen for run 663, but the
final pressure level obtained is similar between the two HEG runs. This is an indication that there is no
change concerning the presence (or absence) of a separation region in this case. This is consistent with
the finding that the pressures in run 663 are already above those which can be explained purely by the
combustion of the fuel.

Figure 3.2.3: Investigation of the influence of fuel injection on the flow in the fuel-on duct separated
from the influence of combustion by using nitrogen as the test gas.

In order to separate the effects of combustion and fuel injection on the flow, and to make an estimate
of the mixing lengths applicable in the scramjet duct an experiment (run 672, Figure 3.2.3) was carried
out in which nitrogen replaced air as the test gas. A baseline pressure from which to calculate the
pressure expected through equilibrium combustion of the fuel could be extracted. In addition, the heat
flux data shows that the cooling effect of the fuel persists until transducers FH04-FH05, indicating a
mixing length of 57-79mm. The increase in wall heat flux after the end of the mixing region is due to
the increased density in the flow after fuel addition.

36
Figure 3.2.4: Comparison of results with and without application of boundary layer trips on the inlet.

In order to test whether the higher pressures in the combustion chamber were caused by separated flow
on the inlet, a boundary layer trip was installed on the intake of the engine. This trip is a 15mm wide
strip of #100 sandpaper and is fixed 20mm from the leading edge of the engine. It is present in run 670
and absent in run 671, and when present should increase the turbulence sufficiently to significantly
affect the size of any separation region on the inlet. Run 670 shows a higher inlet pressure and a lower
inlet wall heat transfer than run 671 indicating a thicker boundary layer caused by earlier transition.
However, no qualitative difference is seen between runs 670 and 671 concerning the measurements
obtained in the combustion chamber.

37
4 Conclusions
The results of the HyShot postflight analysis show that the operating range of HEG has successfully
been extended to allow ground based testing of a complete scramjet engine at M=7.8 in an altitude
range of 27 – 33km. During the available test time, the boundary layers within the HyShot model are
established and stable combustion of hydrogen fuel is observed.

The comparison of the normalised pressure distributions in the fuel-off and the fuel-on combustors
obtained in HEG using air and nitrogen as a test gas and their evaluation by applying analytical and
engineering correlations and the performance of equilibrium combustion calculations strongly
indicates that the flow in the fuel-on duct remains supersonic with combustion. The heat flux data
obtained in HEG provides additional valuable information concerning boundary layer transition from
laminar to turbulent flow.

Pressures on the scramjet intake and in the front of the combustion chamber are consistent with
analytical predictions using oblique shock relations. Pressures in the combustion chamber when no
fuel is present agree with the CFD. Pressure results in the combustion chamber at lower equivalence
ratios agree with predictions of the pressure using an equilibrium combustion model. Pressure results
in the combustion chamber at higher equivalence ratios exceed the predictions of the equilibrium
combustion model and also the predicted pressure at which boundary layer separation should occur,
showing the probable existence of flow separation.

At the HEG condition corresponding to 32.5km flight altitude the boundary layer undergoes natural
transition from laminar to turbulent flow at a Reynolds number of approximately one million in the
fuel-off duct. The heat flux data in the fuel-on duct indicates that the transition process is initiated by
boundary layer tripping due to the fuel injection.

At the lower flight altitude of 27.1km, the flow in the combustor can be regarded as fully turbulent.
The comparison of the data measured in HEG with the flight data of HyShot shows qualitatively good
agreement and confirms the interpretation of the flight data that supersonic combustion was
successfully established. The quantitative discrepancy between flight and ground based test data
increases with decreasing flight altitude.

The parametric studies for the HEG operating condition related to a flight altitude of 27.1km clearly
show the potential of ground based testing in HEG to investigate and verify the flow path development
in complete scramjet engine configurations. Important aspects of the flow path development such as
the laminar-turbulent transition of the boundary layer flow, the development of separation, the
influence of angle of attack, equivalence ratio and dynamic pressure variations and their
interdependencies could be simulated in HEG as part of the HyShot post flight analysis. The data
obtained in HEG suggests that for some flow configurations separation occurs.

Due to the fact that no flow visualization was performed in the present investigations and that detailed
knowledge of the onset of separation is important for the performance prediction of a scramjet
engines, future investigations including flow visualization and accompanying computational fluid
dynamics investigations are required.

38
5 Comments and Recommendations for Future Activities
In the framework of the present investigations, comparisons have been made between the data
obtained in a wind tunnel model and a flight model. CFD comparisons have been made for the fuel-off
case, however, CFD including fuel injection and detailed modelling of the turbulent mixing process,
ignition and combustion has so far not been performed for the HyShot configuration.

It is recommended that future work should include a focus on advanced CFD computations on
the model used in the present study, including fuel injection, mixing, and combustion. In this
way the flight and ground testing data can provide a calibration baseline for CFD codes. Due to
the complex flow structure of the ignition and mixing process, three-dimensional CFD is
required. CFD codes which are suitable to perform these numerical investigations are available
in the ESA member states. However, the detailed modelling of the combustion process is still
very time consuming. In combination with computation of three-dimensional flows, the
generation of one solution requires a timeframe in the order of several month. Therefore,
dedicated research programs are needed for the CFD validation activity.

The deduction of qualitative flow phenomena from pressure and temperature data is possible, but is
often inadequate to have complete confidence in the deductions. The HEG has a number of optical
techniques available including schlieren and interferometry. The utilisation of these flow visualisation
techniques would allow observation of the flow topology in the combustors. Some minor modification
of the model would be required to include glass sidewalls.

It is recommended that further experiments with scramjet models should endeavour to include
optical measurement techniques within the combustion chamber.

Although the use of a simple internal flow path such as used in the HyShot I and II configuration is
useful for basic investigations, more complex, complete models are necessary to approach an
engineering solution to hypersonic airbreathing flight. Thrust measurements need to be undertaken, as
does true coupling of the intake to the combustion chamber.

The next step for experimentation and flight testing should be the investigation of
aerodynamically complete, small-scale engines with the parameter of emphasis being the net
thrust of the complete configuration.

In addition to the problem of net thrust production, the investigation of scaling laws is important. The
ability to take a functioning small engine and create a larger and more powerful engine is vital to the
development of operational scramjet powered vehicles. Scaling is made more difficult by the fact that
viscous effects scale differently to combustion effects.

Therefore, it is recommended to ground test and fly complete scramjet engines at different size
in order to evaluate and validate scaling effects.

In order to realise the reliable and cost-effective access to space the development of new propulsion
systems is required. Air breathing propulsion is one of the technologies which has shown potential in
this framework. This can be demonstrated for example by investigation the potential specific impulse
of rocket and airbreathing propulsion systems shown in Figure 5.1. However, there exist conflicting
studies as to whether scramjets offer high enough efficiencies to be practicable. It is for example
recognized by Heiser & Pratt (1994) that an eventual successful airbreathing engine powered vehicle
will need to employ rocket propulsion for some part of the flight path to orbit, and may indeed
combine rocket and airbreathing propulsion in one engine. Scramjets and rockets are thus extreme
points around a final solution, but it is necessary to understand these extremes in order to correctly
design for example rocket based combined cycle engines (RBCCs). Knowing the separate efficiencies
of scramjet and rocket engines can provide a path to predicting the efficiencies of RBCCs.

39
Figure 5.1: Potential specific impulses for rockets, turbojets ansd airbreathing engines (Kors (1990)).

Currently, working airbreathing technology is implemented up to around Mach 3 (Edwards (2003)) in


the form of turbine and ramjet systems. There are a large number of programs, mainly for military
applications, investigating dual mode ramjet/scramjet engines in the range of Mach 4 to 8. Mach 8 is
an important boundary because it represents the boundary between the regimes where dual-mode
scramjets can be used, and where advanced scramjets are required. If airbreathing engines are to be
used in launch vehicles, it is clear that higher Mach numbers than 8 will need to be investigated (see
for example Falempin & Serre (2003).

It is thus recommended that the next logical range to be investigated in ground based testing and
flight is that between Mach 8 and Mach 12. Currently no research programme including flight
testing is performed in this flight velocity range in the ESA member states.

To clarify whether scramjets have a future as part of new launch vehicles, work is proceeding on
testing the aerothermodynamics and external aerodynamics of scramjets, on looking at systems and
potential problems with future engines, and on testing the viability of the scramjet for space flight.
However, above Mach 8, limited data is available on scramjet performance. Experimentally, this has
been because of a lack of facilities and a lack of instrumentation, and the prohibitive cost of traditional
flight testing. Theoretically, computations have been hampered by a lack of maturity and validation of
CFD codes.

Above Mach 8, the effects of viscous processes on the engine make direct-connect testing of use only
for specific components (such as injectors), as the interactions effects between the intake and
combustion chamber define the operational limits of the engine. Vitiated facilities cannot then be used
for testing in excess of Mach 8. Arc-heated facilities offer some possibilities, but suffer from low
stagnation pressures meaning that simulations of high Mach number flight can only be undertaken at
very high altitudes.

Shock tunnels and instrumentation have matured to provide reliable ground-test facilities for this range
of Mach numbers. It is clear from the HyShot programme (Paull (1999)) and the present study that
scramjet testing in shock tunnels is feasible. The HyShot flight programme was exclusively based on
ground testing in the T4 free piston shock tunnel of The University of Queensland. However, as with
all types of ground based facilities, the test flow includes contaminants whose effects are as yet largely
unquantified. Quantification of these uncertainties requires flight testing. CFD approaches maturity,
but the physical models lack validation. This validation process requires in turn the use of ground
based testing facilities which have been proved to be reliable by comparison with flight tests.

40
It is thus recommended that a programme of coordinated flight tests, ground tests, and CFD
investigation is undertaken so that ground testing facilities can be calibrated using flight test
data and that CFD codes can be validated using a matrix of ground test and flight data.

The performance of scramjet flight tests are at the same time high risk and high pay-off tasks. If
system flight demonstrators or even flight test vehicles aiming at horizontal flight trajectories are used,
flight tests are cost intensive. This is mainly due to the relatively long period of hypersonic flight of 30
s and more. Consequently, the development of the flight test vehicles and the performance of such
flight tests are very complex. Cost effective flight testing has recently become available using
sounding rockets (Figure 5.2), to achieve test times of several seconds. In Australia, two flight test
have so far been undertaken as part of the HyShot flight test programme. The goal of these flights has
been to obtain fundamental data of scramjet combustion in a simple flight configuration at a flight
Mach number of approximately M=8. In the HyShot programme, a number of tests are planned in the
near future employing more complex flow paths and higher flight Mach numbers.

Figure 5.2: Scramjet parabolic flight testing trajectories (Alesi et al. (2003)) based on the
utilisation of sounding rockets.

It is recommended to utilise the flight test concept based on sounding rockets in order to
perform cost effective flight investigations of individual new technologies which are required to
develop a working scramjet powered vehicle with a positive aero-propulsive balance.

41
References
Alesi, H., Paull, A., Paull, R., “HyShot Flight Program: Initial Steps Towards a Scramjet Satellite Launcher“,
presented atASRI Conference, Melbourne, 5 December, 2003.
Anderson, G., Kumar, A., Erdos, J., “Progress in hypersonic combustion technology with computation and
experiment”, AIAA 2nd International Aerospace Planes Conference, 1990, AIAA-90-5254.

Brück, S., Radespiel, R., Longo, J.M.A. “Comparison of nonequilibrium flows past a simplified space-shuttle
configuration”, AIAA 97-0275, 35th Aerospace sciences meeting, January 6-10, 1997, Reno.

Edwards, T. “Liquid fuels and propellants for aerospace propulsion 1903-2003”, J. Propulsion and Power, V19,
No. 6, November-December 2003.

Eitelberg, G. "First results of the calibration and use of the HEG", AIAA 1994-2525, 18th AIAA Aerospace
Ground Testing Conference, Colorado Springs, June 20-23, 1994.

Falempin, F. Serre, L. “LEA flight test program, a first step to operational application of high-speed airbreathing
propulsion”, 12th AIAA International space planes and hypersonic systems and technologies conference,
Norfolk, Virginia 15-19 December 2003.

Gardner, A.D, Jacobs, P.A, Hannemann, K. “End-to-end modelling and design of a new operating condition for
HEG”, in: New Results in Numerical and Experimental Fluid Mechanics IV, Vol. 87, Springer, 2004.

Gardner, A.D., Paull, A., McIntyre, T.J. “Upstream injection in a 2-D Scramjet model”, Shock waves, Vol 11,
No. 5, pp369-375, 2002.

Hannemann, K., Schnieder, M., Reimann, B., Martinez Schramm, J., “The influence and the delay of driver gas
contamination in HEG”, AIAA 2000-2595, 21st AIAA Aerodynamic Measurement Technology and Ground
Testing Conference, Denver, 19-22 June, 2000

Hannemann, K., “High Enthalpy Flows in the HEG Shock Tunnel: Experiment and Numerical Rebuilding”,
AIAA 2003-0978, 41st AIAA Aerospace Sciences Meeting and Exhibit, Reno, 6-9 January, 2003

Heiser, W.H., Pratt, D.T, “Hypersonic Airbreathing Propulsion”, AIAA, Washington, 1994.

Hornung, H.G., Sandeman, R.J., “Interferometric measurements of radiating ionising argon flow over blunt
bodies”, Journal of Physics D, Vol 7, pp920-934.

Huber, P.W., Schexnayer, C.J., McClinton, C.R., “Criteria for self-ignition of supersonic hydrogen-air
mixtures.”, NASA technical paper 1457, 1979.

Korkegi, R.H., “Comparison of shock-induced two- and three-dimensional incipient turbulent separation”, AIAA
journal, Vol 13, No, 4, pp534-535, 1975.

Kors, D.L., “Design considerations for combined air breathing-rocket propulsion systems.” AIAA Paper No. 90-
5216, 1990.

McIntosh, M.K., “Computer Program for the Numerical Calculations of Frozen and
Equilibrium Conditions in Shock Tunnels”, Technical Report, ANU, 1968.

Paull, A., “Hypersonic airbreathing propulsion”, 22nd International symposium on shock waves, 1999.

Paull, A., Alesi, H., Anderson, S., The HyShot Flight Program and how it was developed, AIAA 2002-4939,
AIAA/AAAF 11th International Space Planes and Hypersonic Systems and Technologies Conference,
Orleans, France, 29 September - 4 October, 2002

Schetz, J.A., Billig, F.S., “Penetration of Jets Injected into a Supersonic Stream”,
Journal of Spacecraft, Vol 3, No 11, pp 1658, 1966.

42
A1 Experimental Results for each Individual Run
A1.1 Run 662

Figure A1.1.1: Pressure distributions averaged from 5ms to 7.5ms after shock reflection. Error bars
indicate the scatter in the data.

Figure A1.1.2: Heat flux distributions averaged from 5ms to 7.5ms after shock reflection. Error bars
indicate the scatter in the data.

43
Figure A1.1.3: Comparison between pressure distributions from HEG and from the equivalent
condition in the HyShot trajectory.

44
A1.2 Run 663

Figure A1.2.1: Pressure distributions averaged from 6ms to 9ms after shock reflection. Error bars
indicate the scatter in the data.

Figure A1.2.2: Heat flux distributions averaged from 5ms to 7.5ms after shock reflection. Error bars
indicate the scatter in the data.

45
Figure A1.2.3: Comparison between pressure distributions from HEG and from the equivalent
condition in the HyShot trajectory.

46
A1.3 Run 664

Figure A1.3.1: Pressure distributions averaged from 6ms to 9ms after shock reflection. Error bars
indicate the scatter in the data.

Figure A1.3.2: Heat flux distributions averaged from 6ms to 9ms after shock reflection. Error bars
indicate the scatter in the data.

47
Figure A1.3.3: Comparison between pressure distributions from HEG and from the equivalent
condition in the HyShot trajectory.

48
A1.4 Run 666

Figure A1.4.1: Pressure distributions averaged from 4.5ms to 7ms after shock reflection. Error bars
indicate the scatter in the data.

Figure A1.4.2: Heat flux distributions averaged from 4.5ms to 7ms after shock reflection. Error bars
indicate the scatter in the data.

49
Figure A1.4.3 :Comparison between pressure distributions from HEG and from the equivalent
condition in the HyShot trajectory.

50
A1.5 Run 667

Figure A1.5.1: Pressure distributions averaged from 4.5ms to 7.5ms after shock reflection. Error bars
indicate the scatter in the data.

Figure A1.5.2: Heat flux distributions averaged from 4.5ms to 7.5ms after shock reflection. Error bars
indicate the scatter in the data.

51
Figure A1.5.3: Comparison between pressure distributions from HEG and from the equivalent
condition in the HyShot trajectory.

52
A1.6 Run 668

Figure A1.6.1: Pressure distributions averaged from 4.5ms to 9ms after shock reflection. Error bars
indicate the scatter in the data.

Figure A1.6.2: Heat flux distributions averaged from 4.5ms to 9ms after shock reflection. Error bars
indicate the scatter in the data.

53
Figure A1.6.3: Comparison between pressure distributions from HEG and from the equivalent
condition in the HyShot trajectory

54
A1.7 Run 670

Figure A1.7.1: Pressure distributions averaged from 4.6ms to 7.6ms after shock reflection. Error bars
indicate the scatter in the data.

Figure A1.7.2: Heat flux distributions averaged from 4.6ms to 7.6ms after shock reflection. Error bars
indicate the scatter in the data.

55
Figure A1.7.3: Comparison between pressure distributions from HEG and from the equivalent
condition in the HyShot trajectory averaged from 4.6ms to 7.6ms after shock reflection.

56
Figure A1.7.4: Pressure distributions averaged from 9.5ms to 9.8ms after shock reflection. Error bars
indicate the scatter in the data.

Figure A1.7.5: Heat flux distributions averaged from 9.5ms to 9.8ms after shock reflection. Error bars
indicate the scatter in the data.

57
Figure A1.7.6: Comparison between pressure distributions from HEG and from the equivalent
condition in the HyShot trajectory averaged from 9.5ms to 9.8ms after shock reflection.

58
A1.8 Run 671

Figure A1.8.1: Pressure distributions averaged from 5.0ms to 8.0ms after shock reflection. Error bars
indicate the scatter in the data.

Figure A1.8.2: Heat flux distributions averaged from 5.0ms to 8.0ms after shock reflection. Error bars
indicate the scatter in the data.

59
Figure A1.8.3: Comparison between pressure distributions from HEG and from the equivalent
condition in the HyShot trajectory averaged from 5.0ms to 8.0ms after shock reflection.

60
Figure A1.8.4: Pressure distributions averaged from 9.7ms to 10.0ms after shock reflection. Error
bars indicate the scatter in the data.

Figure A1.8.5: Heat flux distributions averaged from 9.7ms to 10.0ms after shock reflection. Error
bars indicate the scatter in the data.

61
Figure A1.8.6: Comparison between pressure distributions from HEG and from the equivalent
condition in the HyShot trajectory averaged from 9.7ms to 10.0ms after shock reflection.

62
A1.9 Shot 672

Figure A1.9.1: Pressure distributions averaged from 3.5ms to 7.5ms after shock reflection. Error bars
indicate the scatter in the data.

Figure A1.9.2: Heat flux distributions averaged from 3.5ms to 7.5ms after shock reflection. Error bars
indicate the scatter in the data.

63
Figure A1.9.3: Comparison between pressure distributions from HEG and from the equivalent
condition in the HyShot trajectory averaged from 3.5ms to 7.5ms after shock reflection.

64

View publication stats

You might also like