Sie sind auf Seite 1von 254

Faculty of Engineering

MODELLING AND SIMULATION OF


LIQUID ROCKET ENGINE IGNITION
TRANSIENTS

a Dissertation submitted to the Doctoral Committee


of
Tecnologia Aeronautica e Spaziale
in partial fulfilment of the requirements for the
degree of Doctor of Philosophy

Tutor Candidate
Prof. Marcello Onofri Francesco Di Matteo

Co-tutor
Ing. Marco De Rosa

Academic Year 2010-2011


to my family
Contents

Nomenclature x

1. Introduction 1
1.1. Motivation: what is the ignition transient and why are we inter-
ested in it? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2. Key challenges in Rocket Engine start-up . . . . . . . . . . . . . . 2
1.3. Main objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.4. Organization of this Thesis . . . . . . . . . . . . . . . . . . . . . . 5

2. State of the Art 8


2.1. Engine cycles and their start-up and shut-down transients . . . . 8
2.1.1. Gas Generator Engine . . . . . . . . . . . . . . . . . . . . 8
2.1.2. Expander Engine . . . . . . . . . . . . . . . . . . . . . . . 10
2.1.3. Staged Combustion Engine . . . . . . . . . . . . . . . . . 12
2.2. Modelling: review of previous works . . . . . . . . . . . . . . . . . 15

3. ESPSS: European Space Propulsion System Simulation 19


3.1. Fluid Properties Library . . . . . . . . . . . . . . . . . . . . . . . . 20
3.1.1. Perfect Gas properties according to CEA . . . . . . . . . . 21
3.1.2. Perfect Gas interpolated properties . . . . . . . . . . . . . 23
3.1.3. Simplified Liquid interpolated properties . . . . . . . . . . 24
3.1.4. Real Fluids interpolated properties . . . . . . . . . . . . . 25
3.1.5. Perfect gas mixtures . . . . . . . . . . . . . . . . . . . . . 27
3.1.6. Real Fluid - Perfect gas mixtures . . . . . . . . . . . . . . 28
3.2. Fluid Flow 1D Library . . . . . . . . . . . . . . . . . . . . . . . . . 32
3.2.1. Components Classification . . . . . . . . . . . . . . . . . . 33
3.2.2. Junction/Valve . . . . . . . . . . . . . . . . . . . . . . . . . 35

i
Contents

3.2.3. Capacity/Volume . . . . . . . . . . . . . . . . . . . . . . . 37
3.2.4. Tubes/Pipes . . . . . . . . . . . . . . . . . . . . . . . . . . 39
3.3. Turbomachinery Library . . . . . . . . . . . . . . . . . . . . . . . . 43
3.3.1. Pump & Generic Pump . . . . . . . . . . . . . . . . . . . . 45
3.3.2. Turbine & Generic Turbine . . . . . . . . . . . . . . . . . 47
3.4. Combustion Chambers Library . . . . . . . . . . . . . . . . . . . . 49
3.4.1. Injector Cavity . . . . . . . . . . . . . . . . . . . . . . . . 53
3.4.2. Combustor Equilibrium . . . . . . . . . . . . . . . . . . . 53
3.4.3. Combustor rate . . . . . . . . . . . . . . . . . . . . . . . . 59
3.4.4. Nozzle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
3.4.5. Cooling Jacket components . . . . . . . . . . . . . . . . . 70

4. Steady State Library 78


4.1. Components Overview . . . . . . . . . . . . . . . . . . . . . . . . 78
4.2. Ports . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
4.3. The “type” switch . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
4.4. 1-D pipes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
4.5. 0-D components: junctions & valves . . . . . . . . . . . . . . . . . 83
4.6. Combustion Chambers . . . . . . . . . . . . . . . . . . . . . . . . . 84
4.7. Cooling Channels . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
4.8. Turbomachinery . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
4.8.1. Pump . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
4.8.2. Turbine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
4.9. Validation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
4.9.1. Component validations . . . . . . . . . . . . . . . . . . . . 96
4.9.2. Subsystem validations . . . . . . . . . . . . . . . . . . . . 100
4.9.3. Engine cycle designs . . . . . . . . . . . . . . . . . . . . . 105

5. Transient Modelling 114


5.1. Injector Plate model . . . . . . . . . . . . . . . . . . . . . . . . . . 114
5.1.1. Qualitative behaviour . . . . . . . . . . . . . . . . . . . . 118
5.2. Hot Gas side heat transfer coefficient models . . . . . . . . . . . . 120
5.2.1. Models implemented . . . . . . . . . . . . . . . . . . . . . 120
5.2.2. Validation . . . . . . . . . . . . . . . . . . . . . . . . . . . 123

ii
Contents

5.3. Q-2D stratification model for HARCC . . . . . . . . . . . . . . . . 126


5.3.1. Model description . . . . . . . . . . . . . . . . . . . . . . . 126
5.3.2. Numerical validation . . . . . . . . . . . . . . . . . . . . . 131
5.3.3. Experimental validation . . . . . . . . . . . . . . . . . . . 132

6. Integrated Validation: RL-10 design and analysis 142


6.1. Overview of the RL-10A-3-3A rocket engine . . . . . . . . . . . . 143
6.2. Design procedure . . . . . . . . . . . . . . . . . . . . . . . . . . . 147
6.2.1. Turbomachinery modelling . . . . . . . . . . . . . . . . . 147
6.2.2. Thrust chamber and cooling jacket modelling . . . . . . . 152
6.2.3. Lines, valves and manifolds modelling . . . . . . . . . . . 158
6.3. Subsystem simulation: validation at nominal conditions . . . . . . 162
6.4. RL-10 Engine start-up . . . . . . . . . . . . . . . . . . . . . . . . . 165
6.4.1. Description of the start-up sequences . . . . . . . . . . . . 165
6.4.2. Start transient . . . . . . . . . . . . . . . . . . . . . . . . . 167
6.5. RL-10 engine shut-down . . . . . . . . . . . . . . . . . . . . . . . . 175
6.5.1. Description of the shut-down sequence . . . . . . . . . . 175
6.5.2. Shut-down transient . . . . . . . . . . . . . . . . . . . . . 177
6.6. Dynamic Response Analysis . . . . . . . . . . . . . . . . . . . . . 184

7. Conclusions 197

A. Implementation of Up-wind Roe Scheme b


A.1. Governing equations . . . . . . . . . . . . . . . . . . . . . . . . . . b
A.1.1. 4-equation subset . . . . . . . . . . . . . . . . . . . . . . . c
A.2. Numerical concepts . . . . . . . . . . . . . . . . . . . . . . . . . . . d
A.2.1. Roe’s numerical scheme . . . . . . . . . . . . . . . . . . . d
A.2.2. Approximate Riemann Solver . . . . . . . . . . . . . . . . e
A.3. Reconstruction method . . . . . . . . . . . . . . . . . . . . . . . . i
A.3.1. Higher order accuracy . . . . . . . . . . . . . . . . . . . . i
A.3.2. Preconditioning . . . . . . . . . . . . . . . . . . . . . . . . o
A.3.3. Variable cross-section . . . . . . . . . . . . . . . . . . . . . p

B. Friction Factor Correlations r

iii
Contents

B.1. Single-Phase Friction Factor Calculation. Function hdc_fric . . . r


B.2. Two-Phase Friction Factor Calculation. Friedel Correlation . . . . r
B.3. Elbow Pressure Loss Function . . . . . . . . . . . . . . . . . . . . . s

C. Film Coefficient Calculation v

iv
List of Figures

2.1. Vulcain 2 schematic [46] . . . . . . . . . . . . . . . . . . . . . . . . 9


2.2. Vinci schematic [45] . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.3. Space Shuttle Main Engine schematic [149] . . . . . . . . . . . . . 13
2.4. Space Shuttle Main Engine start-up sequence [12] . . . . . . . . . 14
2.5. Space Shuttle Main Engine shut-down sequence [12] . . . . . . . 15

3.1. Components in the fluid_flow_1d library . . . . . . . . . . . . . 34


3.2. Pipe discretisation . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
3.3. Components in the turbo_machinery library . . . . . . . . . . . 45
3.4. Components in the comb_chambers library . . . . . . . . . . . . 52
3.5. Cooling jacket wall mesh [43] . . . . . . . . . . . . . . . . . . . . . 71
3.6. Simplified Cooling Jacket wall disposition [43] . . . . . . . . . . . 74
3.7. Channel with relevant areas and surfaces for heat flux calculation 75
3.8. Longitudinal heat fluxes for a segment i . . . . . . . . . . . . . . . 77

4.1. Components in the Steady State library . . . . . . . . . . . . . . . 79


4.2. Cooling jacket channels wall mesh [43] . . . . . . . . . . . . . . . 91
4.3. Schematic of the Pipeline test case. Purple: steady state compo-
nents. Cyan: transient components . . . . . . . . . . . . . . . . . . 97
4.4. Schematic of Combustion Chamber test case. Purple: steady state
components. Cyan: transient components . . . . . . . . . . . . . . 99
4.5. Turbopump test case: HM7B power pack transient schematic . . . 102
4.6. Turbopump test case: HM7B power pack steady state schematic . 102
4.7. Chamber test case: HM7B Combustion Chamber transient schematic 103
4.8. Chamber test case: HM7B Combustion Chamber steady state
schematic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
4.9. HM7B engine system schematic . . . . . . . . . . . . . . . . . . . 106

v
List of Figures

4.10. Schematic of the RL-10 engine . . . . . . . . . . . . . . . . . . . . 111

5.1. Schematic illustration of an arbitrary injector head . . . . . . . . 115


5.2. Schematics of the injector plates . . . . . . . . . . . . . . . . . . . 116
5.3. Temperature profiles from original and new model . . . . . . . . . 119
5.4. Heat fluxes and wall temperatures results . . . . . . . . . . . . . . 124
5.5. left: 1-D fluid element and energy balance used for conventional 1-
D method; right: control volumes of the Q-2D approach integrated
in 3D wall elements . . . . . . . . . . . . . . . . . . . . . . . . . . 128
5.6. Cooling jacket wall mesh . . . . . . . . . . . . . . . . . . . . . . . 130
5.7. Methane bulk variables evolution along channel axis . . . . . . . 133
5.8. Design of the 4 sector HARCC segment . . . . . . . . . . . . . . . 134
5.9. Schematic of the experimental test case . . . . . . . . . . . . . . . 136
5.10. Wall and fluid thermal stratification,AR = 9.2, pc = 88 bar . . . . 138
5.11. Wall and fluid thermal stratification, AR = 9.2, pc = 58 bar . . . . 139
5.12. Wall and fluid thermal stratification, AR = 30, pc = 88 bar . . . . 140
5.13. Wall and fluid thermal stratification, AR = 30, pc = 58 bar . . . . 141

6.1. RL-10A-3-3A engine schematic [115] . . . . . . . . . . . . . . . . . 145


6.2. RL-10A-3-3A engine diagram . . . . . . . . . . . . . . . . . . . . . 147
6.3. Pumps performance maps . . . . . . . . . . . . . . . . . . . . . . . 149
6.4. Iterative procedure for determining pump parameters . . . . . . . 150
6.5. Turbine performance maps from P&W [15] . . . . . . . . . . . . . 151
6.6. Turbine performance maps . . . . . . . . . . . . . . . . . . . . . . 152
6.7. RL-10A-3-3A chamber contour [15] and discretisation . . . . . . . 154
6.8. Cooling jacket channels profiles . . . . . . . . . . . . . . . . . . . 157
6.9. Venturi nozzle profile . . . . . . . . . . . . . . . . . . . . . . . . . 159
6.10. RL-10A-3-3A schematic model . . . . . . . . . . . . . . . . . . . . 164
6.11. RL-10A-3-3A Valve schedule for Start-up Simulation [15] . . . . . 166
6.12. Valves opening sequence adopted in the simulation . . . . . . . . 168
6.13. Transient results - part 1 . . . . . . . . . . . . . . . . . . . . . . . . 171
6.14. Transient results - part 2 . . . . . . . . . . . . . . . . . . . . . . . . 172
6.15. Transient results - part 3 . . . . . . . . . . . . . . . . . . . . . . . . 173
6.16. Transient results - part 4 . . . . . . . . . . . . . . . . . . . . . . . . 174

vi
List of Figures

6.17. RL-10A-3-3A Valve schedule for Shut-down Simulation [15] . . . 176


6.18. Valves closing sequence adopted in the simulation . . . . . . . . . 177
6.19. Shut-down results - part 1 . . . . . . . . . . . . . . . . . . . . . . . 180
6.20. Shut-down results - part 2 . . . . . . . . . . . . . . . . . . . . . . . 181
6.21. Shut-down results - part 3 . . . . . . . . . . . . . . . . . . . . . . . 182
6.22. Shut-down results - part 4 . . . . . . . . . . . . . . . . . . . . . . . 183
6.23. TCV throttle results - part 1 . . . . . . . . . . . . . . . . . . . . . . 187
6.24. TCV throttle results - part 2 . . . . . . . . . . . . . . . . . . . . . . 188
6.25. TCV throttle results - part 3 . . . . . . . . . . . . . . . . . . . . . . 189
6.26. TCV throttle results - part 4 . . . . . . . . . . . . . . . . . . . . . . 190
6.27. OCV throttle results - part 1 . . . . . . . . . . . . . . . . . . . . . . 193
6.28. OCV throttle results - part 2 . . . . . . . . . . . . . . . . . . . . . . 194
6.29. OCV throttle results - part 3 . . . . . . . . . . . . . . . . . . . . . . 195
6.30. OCV throttle results - part 4 . . . . . . . . . . . . . . . . . . . . . . 196

A.1. Piece-wise linear reconstruction. . . . . . . . . . . . . . . . . . . . l

B.1. Elbow pressure loss parameters . . . . . . . . . . . . . . . . . . . . t

vii
List of Tables

4.1. 1-D pipe element . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82


4.2. 0-D Junction element . . . . . . . . . . . . . . . . . . . . . . . . . . 83
4.3. Combustor element . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
4.4. Nozzle element . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
4.5. Cooling jacket element . . . . . . . . . . . . . . . . . . . . . . . . . 89
4.6. Regenerative circuit element . . . . . . . . . . . . . . . . . . . . . 90
4.7. Pump element . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
4.8. Turbine element . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
4.9. Pipeline input data . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
4.10. Pipeline output data . . . . . . . . . . . . . . . . . . . . . . . . . . 98
4.11. CC input data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
4.12. CC output data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
4.13. HM7B Turbopump input [44] and initial data . . . . . . . . . . . . 101
4.14. HM7B Turbopump output data . . . . . . . . . . . . . . . . . . . . 101
4.15. HM7B CC input [44] and initial data . . . . . . . . . . . . . . . . . 104
4.16. HM7B CC output data . . . . . . . . . . . . . . . . . . . . . . . . . 104
4.17. HM7B input [44] and initial data . . . . . . . . . . . . . . . . . . . 107
4.18. HM7B engine system output data . . . . . . . . . . . . . . . . . . 108
4.19. RL-10A-3-3A input and initial data . . . . . . . . . . . . . . . . . . 110
4.20. RL-10A-3-3A engine system output data . . . . . . . . . . . . . . . 112

5.1. Injector plate variables comparison . . . . . . . . . . . . . . . . . . 119


5.2. Cooling channels geometries . . . . . . . . . . . . . . . . . . . . . 134
5.3. Positioning of themocouples . . . . . . . . . . . . . . . . . . . . . 135

6.1. RL-10A-3-3A construction data [15] . . . . . . . . . . . . . . . . . 146


6.2. Venturi geometrical data . . . . . . . . . . . . . . . . . . . . . . . . 159

viii
List of Tables

6.3. Fuel line valves parameters . . . . . . . . . . . . . . . . . . . . . . 160


6.4. Oxidiser line valves parameters . . . . . . . . . . . . . . . . . . . . 161
6.5. RL-10A-3-3A engine system output data . . . . . . . . . . . . . . . 163
6.6. Engine dynamic response to TCV ±10% operation . . . . . . . . . 186
6.7. Engine dynamic response to OCV ±10% operation . . . . . . . . . 192

A.1. Different values of ω . . . . . . . . . . . . . . . . . . . . . . . . . . . k

ix
Nomenclature

A Cross section area nst Pump number of stages


AR Aspect ratio P Pressure
C Capacitive P Perimeter
C+ Reduced torque Pr Prandtl number
Cp Specific heat Q Volumetric flow rate
D Diameter Q+ Mass flow coefficient
E Internal energy Q̇ Heat flux
fr Friction factor q̇ Heat flux per unit area
G Mass flow per unit area R, r Geometrical radius
h, H Enthalpy R Universal gas constant
h Dimensionless characteristic head Re Reynolds number
hc Heat transfer coefficient S Entropy
Isp Specific impulse St Stanton number
K Loss coefficient for design condition T Temperature
k Concentrated load losses TDH Total dynamic head
L Characteristic length t Wall thickness
M Mass u Internal energy
M Mach number u Primitive variables vector
MR Mixture ratio V Volume
MW Molar weight v Velocity
mh Enthalpy flow W Turbopump power
N Speed coefficient We Weber number
Nk Number of moles x Fluid quality
Ns Specific speed x Horizontal abscissa
n Reduced speed xnc Non-condensable mass fraction
ns Pump number of suctions Z Compressibility factor

x
Nomenclature

Greek symbols c, cc Combustion chamber


α Fluid void fraction ch Channel
β Volumetric expansivity cond Conductive
β Dimensionless characteristics torque cap Capacitive
γ Variable isentropic coefficient conv Convective
δ total to static pressure ratio cav Cavity
ε Absolute roughness crit Critical condition
ζ Concentrated pressure drop parameter chem Chemical composition vector
η Efficiency eff Effective
ϑ Pump dimensionless parameter ext External
θ Total to static temperature ratio eq Equilibrium condition
κ Isothermal compressibility fu Fuel
λ Thermal conductivity fr Frozen condition
µ Fluid viscosity g, gas Gaseous phase
ν Specific volume gg Gas generator
ν Reduced flow parameter hg Hot gas side
ξ Pipe pressure drop coefficient i, j, k Spatial indices
Π Total to total pressure ratio i Internal wall condition
ρ Density jun Junction
σ Stefan-Boltzmann constant l, liq Liquid phase
τ Mechanical torque mix Mixing condition
τ Time constant nc Non-condensable fraction
τxy Fluid shear stress o Initial condition
φ Interaction parameter ox Oxidiser
φ+ Mass flow coefficient p Pump
ψ+ Head rise coefficient pw Powder
Ω Acentric factor R Rated condition
ω Rotational speed rad Radiative
Subscripts ref Reference point
aw Adiabatic wall sat Saturated condition
amb Ambient sound Sound condition
bu Burned condition t Turbine
c Fluid critical point t Turbulent

xi
Nomenclature

th Throat wet Wet


v, vap Vapour
w Wall

xii
1. Introduction

In this introductory chapter the problem of the ignition start-up and shut-down
transients for liquid rocket engine is discussed. The object of this thesis is not
limited to the modelling and simulation of liquid rocket engines start-up but it
aims to the creation of a tool able to model the dynamic behaviour and to obtain
the functional design of liquid rocket engine systems.
Once the motivation and the interest of this problem are shown, the attention
will be focused on the critical aspects that characterize liquid rocket engine
propulsion system during these particular phases. Finally, the main objectives of
this dissertation as well as a brief overview of its structure will be shown.

1.1. Motivation: what is the ignition transient and why are we


interested in it?

In the past, until the end of the 70s, the starting process development of liquid-
propellant engines was usually achieved empirically by testing different schemes
and start cyclogrammes directly during test firing. It required a big amount of
resources and was time consuming. Start-up calculation methods at that time did
not reflect the main factors affecting the process and could not serve as reliable
means for start development. More than 30% of the engine failures occurred
during start-up [68]. Recently, with more powerful and complex engines, the
need of more rationale and reliable methods of the starting process development
appeared.
The prediction of the start-up characteristics of liquid propellant rocket engines
is important to the engine configuration and control system design processes.
Despite that, engine start systems have received secondary considerations since
high-power performance was the chief design objective.

1
1. Introduction

A careful synchronization of control actions with transient start processes


is required to deliver the smooth and reliable thrust build-up characteristics
desired. For turbopump-fed engines, the ability to power turbomachinery prior to
combustion chamber ignition is an important design concern [7]. Indeed, thrust
build-up can be delayed or inhibited if turbine power is insufficient to accelerate
propellant pumps.
Many difficulties associated with engine start predictions stem from the non-
linear mass flow and heat transfer characteristics associated with filling uncondi-
tioned engine systems with cryogenic propellants.
The use of cryogenic fluids imposes additional problems during start-up when
they flow into a system with ambient wall temperatures. Because of strong
evaporation inside tubes, cryogenic fuels may lead to severe over-pressures and
even Thermal choking during the cool-down processes. Accurate predictions are
especially important in engine cycles where this initial propellant flow provides
all of the available turbine power for starting, e.g., expander cycles.
During the start of liquid-propellant engine the main role belongs to the hy-
drodynamic processes (mass flow variations, pressure surges, phase change, etc.).
Hydrodynamic processes are practically the only ones that can be used to act
upon all other phases of the physical-chemical process forming the start-up of the
engine.

1.2. Key challenges in Rocket Engine start-up

Ever since the first liquid rocket engines were developed, performance analyses
have been implemented to examine their operation under various conditions.
Steady state assessments for design conditions are wide-spread and useful in
pre-design phases [17, 47, 61, 94, 100]. Further testing of the nominal operational
modes are performed during initial testing of new liquid rocket engines when
various tests are run in order to verify that the engine has been constructed
according to the original design and performance requirements. This leads to
engines that are extensively reliable during their nominal operation.
Nominal conditions however do not illustrate the extreme conditions in which
most engine components are required to work during transient phases, such as

2
1. Introduction

start-up, shut-down, and throttling. Transient phenomena range from combustion


high frequency instabilities to water hammer effects in feed lines. High pressure
and temperature peaks, inherent to transient phases, may lead to failures in part
of the engine system. The potentially resulting system failures may cause loss
of payload, serious damage to the ground segment, if not loss of human life.
Anomalies in the past, such as the RL-10 anomaly of the Atlas-Centaur flight
AC-71 in 1992 [15], or the Aestus anomaly of Ariane 5G flight 142 in 2001 [74, 82]
have demonstrated that a hard transient may indeed lead to significant system
failures.
In order to have a better understanding of the main problems that may occur
during start-up and shut-down, a description of the main failures for the Japanese
LE-7 engine and for the American SSME engine are here described.

The Space Shuttle Main Engine (SSME) is the high performance LOX/LH2 engine
which was used in the Space Shuttle, producing a thrust of about 1700 kN by
means of a staged combustion cycle. For this very complex engine 5 years of
analysis were necessary to model the transient behaviour of the propellants and
of the hardware during start-up and shut-down.
The SSME engine was sensitive to small changes to propellant conditions and
valve tuning was critical, a 2% error in the valve position or a 0.1 s timing error
could lead to significant damage to the engine.
A step-by-step approach was necessary to explore the start-up sequence with
small time increments: that required 19 tests, 23 weeks, 8 turbopump replacements
to reach 2 s into 5 s of start-up. Additional 18 tests, 12 weeks and 5 turbo-pump
replacements were necessary to touch the minimum power level [12].

The LE-7 is a high performance LOX/LH2 engine employed in the H-II rocket,
which produces a thrust of about 1000 kN by means of a staged combustion cycle
similar to that of the SSME. During H-II Flight 8 in 1999 a failure occurred in
the LE-7 engine. The failure was determined by mechanical vibration problems
into the fuel pump that caused high cycle fatigue and so the premature engine
failure [129]. During development tests of the LE-7 other issues can be addressed
to phenomena occurring during transient phases [50]:

3
1. Introduction

• High sensitivity to heat transfer from hardware to fluids and to ignition


timing reproducibility of both combustion devices

• Functional instability to inlet pump vibration of pressure

• Problem of rotating cavitation on oxidiser and fuel pumps

• Over power during start-up caused damages to hardware

• Explosion at Main Oxidiser Valve opening

1.3. Main objectives

Given the increased power of today’s computers, more frequent use is made of
CFD codes to perform detailed assessments of the flow behaviour in single engine
components. The goal is to understand observed variations in fluid properties and
to find the source of unexpected behaviours. Such CFD methods however require
extensive computational times, making speedy parametric analyses impossible.
Additionally, most require long input preparation times. Although the qualities
of a thorough and in full-depth analysis may be desirable in the study of specific
single components and scenarios, for complex, multi-component systems, the long
computational times become insurmountable.
An intelligent simplification of the underlying processes allows the reduction
of the 3-D governing partial differential equations to one-dimensional or quasi
1-D differential equations which no longer require complex solution methods thus
allowing much faster computational times. Results obtained from such studies
rely significantly on imposed initial and boundary conditions. Simplifications
introduced in these models often lead to an incorrect transient behaviour which
does not correspond to the measured and observed physics. This is, amongst
other reasons, due to the ignored interaction between downstream and upstream
lying components. Tools concentrating on one component only are thus not
sufficient as they do not help in understanding how components affect each other
during transient phases, what their impact on system frequencies is, and how this
interaction may lead to a major component or system failure.

4
1. Introduction

A system approach is then necessary to take into account all the interactions
between all the components of an engine. This choice is fundamental if a detailed
estimation of the engine transient behaviour is the task.
To this purpose the present work was initiated with the aim of studying and
modelling by numerical tool the ignition transient of liquid rocket engines, improv-
ing and implementing more complex and accurate models in system modelling
tools for transient analysis. This aim is achieved in three steps:

• A suitable physical and mathematical collection of models able to design


and analyse the steady state behaviour of liquid rocket engine systems is
developed and implemented in a numerical code. The model library has
been successfully validated with respect to open literature data, performing
design simulations and off-design analyses of actual rocket engines.

• Basic and simple models used to simulate the injector plate of the engine,
the hot-gas-side heat transfer correlations in the combustion chamber and
the regenerative circuit are exchanged with improved, more sophisticated
and more physical models. These models are able to describe accurately
the convective and radiative heat transfer at the injector plate face, the
hot-gas-side heat transfer coefficient, and to describe also the influence of
thermal stratification in high aspect ratio cooling channels.

• At last, the validation of a design procedure and the models developed


are achieved by the design, the start-up, the shut-down and the dynamic
response simulation of a real engine, the RL-10A-3-3A.

1.4. Organization of this Thesis

The work performed is here presented in six chapters as described below.

Chapter 1 provides a brief introduction of the motivations which led this work,
followed by a description of the main issues related to transient phases in liquid
rocket engines. Issues presented practical consequence as described for the case of
LE-7 engine failures or the SSME start-up campaign.

5
1. Introduction

In addition, a description of the main objectives of this work have been provided
explaining which approach has been adopted for the modelling of the components’
behaviour of a liquid rocket engine, and what is the final aim of such a Ph.D.
dissertation.

Chapter 2 is dedicated to a summary of the state of the art in the field of liquid
rocket engine transient simulation. Models and tools from USA, Europe, Russia,
Japan, Iran, China are collected, studied and compared to each other in order
to find advantages and drawbacks present in each work.The literature study is
fundamental to understand in which direction this work should be directed.

Chapter 3 summarizes the ESPSS library. It describes the models implemented in


the European Space Propulsion System Simulation library, a collection of models
for each engine’s component used as starting point for this work. The chapter
includes the way the library and the main components are modelled: fluid prop-
erties, pipes, volumes, valves and junctions components, pumps and turbines,
combustion devices, cooling channels and nozzles.
The assumptions behind each formulation as well as the way components inter-
act with each other are analysed; advantages and limitations of each models are
presented in this chapter.

Chapter 4 illustrates the steady state modelling. It examines the most important
engine components models developed for the creation of a steady state library.
The main purpose of these components is the design and the analyses in steady
state conditions liquid rocket engine systems cycles.
The development of these models is used to enhance the system capabilities of
the code and create a fundamental instrument to be used along the entire design
period from pre-design phase to parametric studies for fine tuning of the engine
parameters.
The Steady State library presented within this chapter enables to perform iter-
ative engine design loops and parametric studies in a reasonable computational
time.

6
1. Introduction

Chapter 5 concerns the transient modelling. It provides a description of all


the new models developed to enhance and improve the simulation capabilities
during transient phases of a propulsion system. Three new models are here im-
plemented and discussed: a new injector plate model, a new formulation for the
evaluation of the heat fluxes in combustion devices, finally a new model for the
evaluation of thermal stratification on cooling channels is described.
For each model a discussion of the derived equations is provided as well as test
cases for validation purposes. Where available models simulations are compared
with numerical test cases or experimental results.

Chapter 6 describes the system level validation. It contains the main test case
examined: the RL-10A-3-3A rocket engine which presents significant challenges
due to its expander cycle configuration and its ignition settings. Modelling of the
liquid rocket engine in the frame of its transient phases (start-up and shut-down)
and the resulting simulation data are discussed.

Chapter 7 concludes this thesis summarising the main achievements of the work
performed and indicating those points which, in the frame of further work, could
be improved and others which could be considered in a further extension of the
models presented.

7
2. State of the Art

2.1. Engine cycles and their start-up and shut-down transients

An engine cycle for turbopump-fed engines describes the specific propellant flow
paths through the major engine components, the method of providing the hot gas
to one or more turbines, and the method of handling the turbine exhaust gases.
Depending of the propellant path considered we will have (among the main ones)
a gas generator cycle, an expander cycle or a staged combustion cycle.
Each system displays a specific sequence for start-up and shut-down phases.
The sequences are tailored to fit general engine characteristics such as engine
cycle, as well as more specific details such as ignition system type. A gas-generator
engine will therefore have a different sequence to an expander cycle or a staged
combustion engine.

2.1.1. Gas Generator Engine

In the gas generator cycle the turbine inlet gas comes from a separate gas generator.
Its propellants can be supplied from separate propellant tanks or can be bled off
the main propellant feed system. This cycle is relatively simple; the pressures in
the liquid pipes and pumps are relatively low (which reduces inert engine mass).
It has less engine specific impulse than an expander cycle or a staged combustion
cycle. The pressure ratio across the turbine is relatively high, but the turbine or
gas generator flow is small (1 to 4% of total propellant flow) if compared to closed
cycles.
Alternatively, this turbine exhaust can be aspirated into the main flow through
openings in the diverging nozzle section. This gas then protects the walls near the
nozzle exit from high temperatures. Both methods can provide a small amount of
additional thrust. The gas generator mixture ratio is usually fuel rich (in some

8
2. State of the Art

engine it is oxidizer rich) so that the gas temperatures are low enough (typically
900 to 1350 K) to allow the use of uncooled turbine blades and uncooled nozzle
exit segments [133, 94].
The European Vulcain and Vulcain 2 (see Figure 2.1) engines are two examples
of gas generator cycle engines. The gas generator in these engines provides hot
gases to power two turbines, one for each propellant pump.

Figure 2.1.: Vulcain 2 schematic [46]

Prior to commencing the start-up sequence the engines undergo a chill-down


phase of all components with the exception of the main propellant valves and gas
generator, to reduce thermal shock effects when the cold fuel is fed to the engine
at start-up. This phase lasts ca. 2.5 hours. The engines are then started using a
starter to move the turbine in order to provide a minimal amount of power to
the pumps to feed the gas generator. A pyrotechnic igniter is then used to ignite

9
2. State of the Art

the gas generator. Given the now full power available to the turbines to run the
pumps, feeding of the main combustion chamber is possible.
The fuel valves are opened first, the regenerative cooling channels fill with
fuel which then irrupts into the combustion chamber which, with aid of another
pyrotechnic igniter, is ignited under lean conditions. The combination of the fuel
rich condition at ignition and the cooling performed by the regenerative cooling
system increases the complexity of the ignition process of the main combustion
chamber [91].

2.1.2. Expander Engine

In the expander cycle most of the engine coolant (usually the fuel) is fed to low-
pressure-ratio turbines after having passed through the cooling jacket where it
picked up energy. Part of the coolant, perhaps 5 to 15%, bypasses the turbine
and rejoins the turbine exhaust flow before the entire coolant flow is injected
into the engine combustion chamber where it mixes and burns with the oxidizer.
The primary advantages of the expander cycle are good specific impulse, engine
simplicity, and relatively low engine mass. In the expander cycle all the propellants
are fully burned in the engine combustion chamber and expanded efficiently in
the engine exhaust nozzle [133, 94].
The American RL-10 engine and the European Vinci engine (currently under
development) represent two examples for this type of cycle (see Figure 2.2).
The latter is one such expander engine which implements a regenerative cooling
system based on hydrogen to cool the combustion chamber. Hot gases generated
in an electric igniter are implemented as a pilot flame to start the main combustion
chamber in which the temperature progressively increases as stable combustion
is established. The liquid hydrogen from the tanks leaving the cooling system
undergoes a phase change from liquid to supercritical during the ignition sequence
and thus powers up the turbines as it gains velocity and pressure. For these very
reasons, the Vinci engine unlike other European engines implements reaction
rather than impulsive turbines [91, 38].
The use of LH2 and LOX as propellants in the RL-10 engine, makes necessary a
chill-down step to avoid propellant boiling and pump cavitation during operation.
Cavitation must be avoided due to its propensity to cause erosion damage to the

10
2. State of the Art

Figure 2.2.: Vinci schematic [45]

blades. When the propellants are used for chill-down, the pre-start valves are
opened to allow flow through the pumps and the tank head pressure provides the
driving force. The LOX is sent through its entire circuit and fed into the main
chamber. The LH2 , however, is flowed out through the cooldown valves prior to
going through the regenerative cooling tubes. This will allow the cooling tubes to
remain at a significantly higher temperature for start-up and avoid any possibility
of inadvertent propellant mixing and igniting prior to start-up.
When the engine start command is given, the cooldown valves are closed and
the main fuel valve is opened. This allows for the LH2 to flow through the
regenerative cooling lines, which heat up and vaporise the fuel. Once the gaseous
hydrogen begins to drive the turbine, the pumps take over providing the pressure
gradient for flow. The main chamber is ignited using a pilot flame created from
gaseous propellants tapped off from the main lines.
One the combustion chamber is lit, additional heat is transferred into hydrogen
flow form combustion process. This drives the turbine to its nominal power and
eventually the engine reaches a steady state condition.
When the shut-down command is issued, the main fuel shut-off valve is closed

11
2. State of the Art

tu cut-off fuel flow into the main chamber. As a result, combustion in the combus-
tion chamber and thrust are rapidly terminated. The prestart valves are closed
to stop the propellant flow and the cooldown valves are opened to prevent over
pressure of the engine due to trapped hydrogen.

2.1.3. Staged Combustion Engine

In the staged combustion cycle, the coolant flow path through the cooling jacket
is the same as that of the expander cycle. Here a high-pressure precombustor
(gas generator) burns all the fuel with part of the oxidizer to provide high-energy
gas to the turbines. The total turbine exhaust gas flow is injected into the main
combustion chamber where it burns with the remaining oxidizer. This cycle lends
itself to high-chamber-pressure operation, which allows a small thrust chamber
size. The extra pressure drop in the precombustor and turbines causes the pump
discharge pressures of both the fuel and the oxidizer to be higher than with
open cycles, requiring heavier and more complex pumps, turbines, and piping.
The turbine flow is relatively high and the turbine pressure drop is low, when
compared to an open cycle. The staged combustion cycle gives the highest specific
impulse, but it is more complex and heavy [133, 94].
A variation of the staged combustion cycle is used in the Space Shuttle Main
Engine (SSME) (see schematic in Figure 2.3).
The engine assembly consists of the engine powerhead and main combustion
chamber/nozzle assembly. The powerhead uses two preburners, a main injector,
and an oxidizer heat exchanger, all welded into the hot gas manifold. The preburn-
ers generate fuel-rich combustion gases to drive the LOX and LH2 turbopumps.
Hydrogen fuel is also used to cool the main chamber and nozzle.
Prior to starting the SSME engine, the start preparation phase takes place. This
consists of purging and thermal conditioning followed again by purging. During
the first purging phase, dry nitrogen and dry helium are used to remove moisture
as well as air which would otherwise freeze along the oxidiser and fuel lines
respectively. After thorough purging has been performed, thermal conditioning is
undertaken by allowing propellants to flow into the engine down to the main fuel
valve (MFV) on the LH2 side and down to three oxidiser valves on the LOX side,
i.e. the main oxidiser valve (MOV), the oxidiser preburner valve (OPOV) and fuel

12
2. State of the Art

Figure 2.3.: Space Shuttle Main Engine schematic [149]

preburner oxidiser valve (FPOV). LH2 and LOX recirculation flows, through bleed
valves, are maintained for about one hour to chill the four turbopumps down to
cryogenic temperatures and to eliminate gas pockets in the propellant feed system.
Once thermal conditioning is completed, a final dry helium purging of the fuel
lines downstream the MFV is performed [78, 12]. At t = 0 s the start command is
sent and the MFV is opened completely and the three spark igniters are provided
with electrical power. The three main oxidiser valves, MOV, FPOV, and OPOV are
then regulated in order to reach the target priming times for the new preburners
and main combustion chamber (MCC).
These are precisely selected to ensure a stable ignition process and are at 1.4 s
for the fuel preburner (FPB), 1.5 s for the MCC, and 1.6 s for the oxidiser preburner
(OPB). Pressure oscillations in the fuel system, arising form thermodynamic
instabilities during the expander-cycle-like start-up, are closely monitored and
the FPOV and OPOV positions are controlled to avoid high mixture ratios in the
preburners which result in dangerously high temperatures for the turbines.
The FPB is ignited during the second fuel system pressure dip and is followed
by ignition of the PFB. At 1.25 s the rotational speed of the high pressure fuel

13
2. State of the Art

Figure 2.4.: Space Shuttle Main Engine start-up sequence [12]

turbopump is checked to ensure that hydrogen can be bumped through the system
against the back-pressure created by the MCC oxidiser priming. As the drive
power of both high pressure turbines increases, the chamber coolant valve (CCV)
is throttled down to 70%. This conditon is maintained until 2.4 s at which point
the control system measures the MCC pressure and regulates the OPOV, FPOV,
and CCV in order to follow the pre-programmed chamber pressure ramp until
the nominal operational point has been reached. Finally the FPOV is regulated to
adjust the fuel mass flow rate until the nominaol mixture ratio is obtained. At 5 s
stable operation has been achieved.

During the shut-down phase (see Figure 2.5), the main goal of which is to
ensure a safe and as quick as possible shut-down of the engine, the OPOV is
the first valve to be closed with a closing rate not higher than 45%/s to avoid a
too abrupt thrust decay which would endanger the orbiter’s structural integrity.
Closing of the FPOV follows. Positioning of both valves is monitored to main-
tain a low mixture ratio and maximum oxidiser pressure decay whilst avoiding

14
2. State of the Art

Figure 2.5.: Space Shuttle Main Engine shut-down sequence [12]

back-flow of hot gases into oxidiser lines. In order to compensate for the increased
heat loads due to throttling, the chamber coolant valve is regulated to force more
coolant into the main combustion chamber. Simulataneously the MOC is closed
at a control rate to ensure a combusiton chamber pressure above the inlet turbine
pressures. Finally after 1 s of additional MFV opening time to assure a very
fuel-rich shut-down, the MFV and the CCV are closed [12, 91].

2.2. Modelling: review of previous works

The development of software tools for analyses of rocket engine systems is critical
to the successful design and fine tuning of such systems. Typically, the ignition
process of a liquid rocket engine involves non-linear interactions between multiple
engine components with phenomena such as flow resistance, off-design turbop-
ump operation, heat transfer, phase change, and combustion. Furthermore, the
physical properties of liquid propellants and combustion products in such systems
vary widely and in a rapid way. Developing tools for predicting the dynamical

15
2. State of the Art

behaviour of an engine with such characteristics is a challenging but important


task for engineers and researchers.

Worldwide, various tools have been developed to simulate the transient behaviour
of rocket engine systems. In 1990, Ruth et al. [124] developed the Liquid Rocket
Transient Code (LRTC) the in-house code of The Aerospace Corporation. It
represents one of the first attempts to simulate propulsion systems with a modular
structure. LRTC models an engine through a modular scheme with the method
of characteristics for a flow through line segments (pipes) connected by nodes
(zero-dimensional components such as valves, orifices, pumps, branches etc.).
Comparisons with Titan IV K-01 flight data of the Stage I start transient demon-
strated general agreement.

The Rocket Engine Transient Simulator (ROCETS) [13] was designed and de-
veloped during the 90s by Pratt & Whitney for NASA-MSFC; it allows for cost-
effective computer predictions of liquid rocket engine transient performance. The
most popular application of ROCETS is the RL-10A-3-3A rocket engine [13, 14, 15],
varying from start transient analysis to modelling of thrust increase with densified
propellants [59, 58].
Another powerful tool created in the United States is the Generalized Fluid
System Simulation Program (GFSSP) for modelling cryogenic fluids in a complex
flow circuit [80].
Recently, other researchers and engineers have developed (or have started to)
other codes, for the transient analysis of propulsion systems, but their work was
mainly focused to only a part of an engine system [7, 22, 31].

In Japan, during the 90s, a quasi-steady simulation code for transient analy-
sis of the original LE-7 engine was developed [70]. This was Japan’s first attempt
to develop a staged combustion cycle engine, and establishing a safe and reliable
start-up and shut-down method was very important.
Later in 2002, the Visual Integrated Simulator for Rocket Engine Cycle (VIS-
REC) [6] was developed by Mitsubishi Heavy Industries. VISREC is a one dimen-
sional flow and heat analysis program using the lumped parameter approach. It

16
2. State of the Art

can analyse start-up and shut-down transient behaviour of many types of engine
cycles such as expander, staged combustion, and gas generator. Together with
the LE-7A rocket engine, the Rocket Engine Dynamic Simulator (REDS) [152] in
2004 was developed and applied to start-up and shut-down transient analyses.

In China, in 2000, Kun et al. [77] have developed a tool to study rocket en-
gine system transient based on the disassembly method whose principles are the
following: the modules have independent physical function and mathematical
model; there are uniform parameters exchange interfaces between each module,
and finally, the engine system can be disassembled into modules by practical
physical units.

Also in India, during the 2000s, a first try to develop a “dynamic simulator
for liquid-propellant rocket engines” has been accomplished; this tool is called
CRESP-LP [134].

In Iran, only recently they have transient simulation tools performed by the
University of Technology in Tehran. Prof. Karimi et al. have developed their
own code and performed several analyses to study transient regime in rocket
engines [71, 72, 73, 119]

In Russia, extensive studies have been performed on engine transient behaviour.


These studies have taken advantage of the extensive database generated during
testing of the wide range of liquid rocket engines that Russia has developed. The
problem of transient phase in liquid rocket engine systems has been studied in
[68, 136]. The transient analysis of liquid rocket engine is illustrated in great detail
in [10]. A simple ordinary differential equation (ODE) approach is presented and
complemented with experimental-based empirical equations in those cases where
ODEs are not sufficient to describe the occurring phenomena.

Finally in Europe, CNES tested a dedicated library for the modelling and simula-
tion of rocket engine system dynamics developed in the AMESim platform [121].
The platform carries multiple sub-systems of a rocket engine, such as tanks, pneu-

17
2. State of the Art

matic lines, turbopump, regenerative circuit, combustion chamber, and starters.


In cooperation with ONERA, CNES has also developed another tool called
Carins [102, 84], an open platform featuring the “symbolic manipulation” method
to simulate the transient behaviour of propulsion systems.
The Astrium SMART code has been developed for the EPS start-up simula-
tion [74] and recently, from 2008 a complete set of models able to simulate liquid
propulsion system components called European Space Propulsion System Simu-
lation (ESPSS) has been developed by a joint European team in the frame of a
GSTP Programme for the European Space Agency [32].

18
3. ESPSS: European Space Propulsion
System Simulation

This chapter documents the models of the propulsion system library implemented
within the existing analysis software EcosimPro [40], used as a basis for this Ph.D.
research.
EcosimPro is an object-oriented visual simulation tool capable of modelling
various kinds of dynamic systems represented by differential-algebraic equations
(DAE) [18] or ordinary-differential equations (ODE) and discrete events. The
modelling of physical components is based on the EcosimPro language (EL), an
object-oriented programming language which is very similar to other conventional
programming languages but is powerful enough to model continuous and discrete
processes. It can be used to study both stationary states and transients.
EcosimPro employs a set of libraries containing various types of components
which can be interconnected to model complex dynamic systems:

• control

• math

• mechanical

• ports_lib

• thermal

The European Space Propulsion System Simulation (ESPSS) consists of mul-


tiple libraries to represent a functional propulsion system, e.g. fluid properties,
pipe networking including multi-phase fluid flow, two-phase two fluids tanks,
non-adiabatic combustion chambers, chemistry, turbomachinery, etc:

19
3. ESPSS: European Space Propulsion System Simulation

• fluid_properties

• fluid_flow_1d

• comb_chambers

• tanks

• turbo_machinery

The Libraries sections hereafter describe those libraries [43], focussing on their
physical modelling, and on the main models used for this Ph.D thesis.

3.1. Fluid Properties Library

fluid_properties is an EcosimPro library in charge of the calculation of fluid


properties. The functions available on this library are mainly used by the
fluid_flow_1d library for the simulation of fluid systems. The most impor-
tant features are summarized as follows:

• Most of the fluids used for rockets applications are available

• Fluids are supported in different categories depending on the type used:

- Perfect gases (transport and heat capacity properties obtained from


CEA polynomials (temperature dependent)). Only used in the combus-
tor/nozzle components
- Perfect gases (transport and heat capacity properties interpolated from
tables (temperature dependent))
- Simplified liquids interpolated from tables (temperature dependent)
- Real fluids interpolated from tables considering either liquid, super-
heated, supercritical or two-phase flow (temperature and pressure
dependent)
- User-defined fluids are available. The properties must be defined in
external data files and can be of any of the last three types previously
mentioned

20
3. ESPSS: European Space Propulsion System Simulation

• Mixtures of a real fluid with a non-condensable gas are allowed. The


homogeneous equilibrium model is used to calculate the properties (quality,
void fraction, etc) in case of two phase flow. Mixtures of two real fluids are
not allowed. Therefore, phenomena such as fractional distillation are not
modelled.

The fluid_properties library does not contain any component. It only provides
a large collection of functions returning the value of a fluid property (or the com-
plete thermodynamic state) by introducing relevant parameters (i.e. temperature,
internal energy, pressure, density, heat transfer and friction correlations etc).

3.1.1. Perfect Gas properties according to CEA

The programming is based on the perfect gas state equation. The expressions used
are summarized as follows:

R·T
P =ρ
MW
Z = 1; β = 1/T ; κ = 1/P

The expressions used for the energy calculation are based on the computation
of the specific heat at constant pressure for ideal gases (Cp0 ) as a function of
temperature only (by means of polynomial expressions). The expression proposed
was obtained from a very large database providing data for a very wide range of

21
3. ESPSS: European Space Propulsion System Simulation

temperatures (between 200 and 20.000 K) and is summarized as follows:

Cp
= a1 T −2 + a2 T −1 + a3 + a4 T + a5 T 2 + a6 T 3 + a7 T 4 (3.1)
R
Z T
H = H(T0 ) + Cp(T )dT =
T0
T2 T3 T 4 b1
 
−2 −1 T
R · T · −a1 T + a2 T ln T + a3 + a4 + a5 + a6 + a7 +
2 3 4 5 T
(3.2)
T
Cp(T ) R P R
Z
S = S(T0 , P0 ) + dT − log( ) = − ·
T0 T MW P0 MW
T2 T3 T 4 b2
  
P −2 −1
log R −a1 T − a2 T + a3 ln T + a4 T + a5 + a6 + a7 +
P0 2 3 4 T
(3.3)

The functions giving the viscosity and the thermal conductivity in case there
are data available (mainly from NIST species database [88], in this case the NIST
database is not as extensive as for the thermodynamic properties) are also based
on polynomial expressions. The viscosity and thermal conductivity functions have
respectively the following form:

B1 C1 B2 C2
ln µ = A1 · ln T + + 2 + D1 ; ln λ = A2 · ln T + + 2 + D2
T T T T

Otherwise the properties are estimated as follows:

Viscosity: There are different estimation methods available. The approach used
is subjected to the availability of property data: critical properties and the dipole
moment. In this case the expression used [120] is:

0.807 Tr0.618 − 0.357 exp(−0.449 Tr ) + 0.340 exp(−4.058 Tr ) + 0.018 · FPo · FQo


 
µ=
ζ · 107

where Tr is the reduced temperature computed as follows Tr = T /Tc , and FQo are
correction factors and ζ is the reduced inverse viscosity, calculated as follows:
 1/6
Tc
ζ = 0.176 ·
MW 3 Pc4

22
3. ESPSS: European Space Propulsion System Simulation

FPo is a correction factor that mainly depends on the polarity of the molecule. It
is computed as follows:

FPo = 1 0 ≤ µr ≤ 0.022
FPo = 1 + 30.55 · (0.292 − Zc )1.72 0.022 ≤ µr ≤ 0.075
FPo = 1 + 30.55 · (0.292 − Zc )1.72 |0.96 + 0.1 · (Tr − 0.7)| 0.075 ≤ µr

FQo is a correction factor used only in quantum gases. In the present case its
value is 1. µr is the relative dipole moment computed as follows:

µ 2 Pc
µr = 52.46 ·
Tc2

where µ is the dipole moment in Debyes.


In case critical properties and dipole moment are not available, other estimations
must be done based on quantum formulation. The following expression will be
used: √
ηns · MWk · T
µk =
Ωk · 107
where ηns is a constant (26.6958 in S.I.) and Ωk

50 · MWk4.6
 
Ωk = ln
T 1.4

Thermal Conductivity: Similarly to viscosity, the thermal conductivities that


are not available are estimated. The approach used is summarized as follows:

µk R (3.75 + 1.32 · Cpk /R − 2.5)
λk =
MWk

3.1.2. Perfect Gas interpolated properties

For the state equation, the same expressions as in Chapter 3.1.1 are used. The
expressions used for the energy calculation are based on the table interpolation of

23
3. ESPSS: European Space Propulsion System Simulation

the specific heat at constant pressure (Cp0 ) as a function of temperature only:

Cp Cp
= (T )
R R
Z T
H = H(T0 ) + Cp (T )dT
T0
T
Cp (T ) R
Z
S = S(T0 , P0 ) + dT − log(P/P0 )
T0 T MW

The functions giving the viscosity and the thermal conductivity are also inter-
polated from the external user-defined property file as a function of temperature.

3.1.3. Simplified Liquid interpolated properties

For simplified liquids, the formulation is based on the tables where, density, sound
speed and specific heat are interpolated as functions of the temperature. Thus, the
volumetric expansivity can be obtained as follows:

−1 dρ
β=
ρ dT P =const

This derivative is calculated numerically. Then, the following others thermody-


namic derivatives can be calculated:

Cp
Cv = 2
1 + T · β 2 νsound /Cp
1 dρ
Cp
κ= = 2
ρ dP T =cte ρ νsound Cv

Once these properties have been interpolated, the equation of state can be
applied assuming constant compressibility with pressure:

 
ρ(P, T ) = ρ(T ) 1 + κ (P − Pref )

The enthalpy is calculated integrating numerically the Cp:


Z T
H = Hideal (T0 ) + Cpideal (T )dT
T0

24
3. ESPSS: European Space Propulsion System Simulation

Viscosity and Thermal conductivity are also interpolated from the external data
of the property file as a function of temperature.

3.1.4. Real Fluids interpolated properties

For real fluids, FORTRAN functions will perform special searching techniques in
2-D property tables to interpolate the desired property in the nearest cell of the
data tables. A single reading will be done the first time that a function’s call of
any property is made. The functions will identify if a certain fluid is liquid, vapour
or two-phase flow by giving a pair of variables that can be ρ-T, ρ-U, s-H , P-H ,
P-T, etc.
With the exception of two-phase flow or in the case of table extrapolation, no
special hypothesis concerning the Equation of State and the properties has been
done: all the properties are interpolated using the data tables of the properties file.
Under two phase conditions the quality of the mixture is calculated from
the saturation properties of the liquid and steam phases. Knowing the mixture
(vapour/liquid) density and energy, the following two equations are used:

u − uliq
quality = x =
uvap − uliq
1/ρ = 1/ρliq + x (1/ρvap − 1/ρliq )

An iterative process in pressure is needed. For each iteration, the saturation


conditions will be calculated and the pressure fulfilling the two previous equations
can be found. The void fraction is calculated as follows:

void fraction = α = Vvap /(Vvap + Vliq ) = (ρliq − ρ)/(ρliq − ρvap )

The transport properties and the heat capacity in two-phase conditions are
calculated in a simple way:

µ = x µvap + (1 − x)µliq (3.4)


λ = x λvap + (1 − x)λliq (3.5)
Cp = x Cpvap + (1 − x)Cpliq (3.6)

25
3. ESPSS: European Space Propulsion System Simulation

Nevertheless, the liquid and vapour saturation values of the transport properties,
together with those of the heat capacities, densities and enthalpies (latent heat)
will be returned for the calculation of other two-phase properties as the sound
speed and the film coefficient.
Under one-phase conditions the sound speed is directly given by the FORTRAN
function interpolations. The returned value is equivalent to the following ex-
pression that uses other properties (Cp, β , κ) also returned by the properties
function: s
Cp
vsound = (3.7)
ρκCp − β 2 T

Under two-phase conditions the sound speed must be calculated. The equilib-
rium sound speed presents discontinuities at phase changes. In order to ensure
system robustness the following approach is given by Wallis (1969) [147]:

2 2 2
1/vsound = (αρvap + (1 − α)ρliq )(α/ρvap /vsound,vap + (1 − α)/ρliq /vsound,liq ) (3.8)

Another “frozen” [28, 8] sound speed expression can be used instead, which is
also continuous with the sound speed at phase changes:

dP x Cpv + (1 − x)Cpl
= ν : specific volume; x: quality
dT T (xβv νv + (1 − x)βl νl )
2
dP v

T
2 dT
vsound =
x  Cpv − T dP (1 + ) βv − κv dT + (1 − x)  Cpl − T dP
dP dP
 
dT νv dT ν l (1 + ) βl − κl dT

where:

 = 0 ⇒ “frozen” sound speed


 = 1 ⇒ “equilibrium” sound speed

Sub indexes “v” and “l” indicate vapour and liquid saturated conditions.

26
3. ESPSS: European Space Propulsion System Simulation

3.1.5. Perfect gas mixtures

Perfect gas mixtures are calculated with linear mixing rules assuming the same
temperature for all the constituents.

nchem
ρk R·T X
xk = ; Pk = ρk ; P = Pk
nchem MWk
k=1
X
ρk
k=1

where, R is the gas constant = 8314.4 [J/kmol K], MWk is the molecular weight of
the chemical constituent k, T is the mixture temperature [K], ρk is the density of
the chemical constituent k and xk is the mass fraction of the chemical constituent
k.
The molecular weight of the mixture MWmix and the molar fractions yk are
calculated as follows:
nchem
1 X xk xk MWmix
= ; yk = molar fraction =
MWmix MWk MWk
k=1

The energy properties are computed as follows:

nchem
X nchem
X
Cp = xk Cpk (T ); Cv = xk Cv,k (T );
k=1 k=1
nchem
X nchem
X nchem
X
H= xk Hk (T ); S= xk Sk (T ) + yk ln yk (T )
k=1 k=1 k=1

The sound speed is calculated as follows:

Cp
γ= (3.9)
Cp − R/MWmix
p
vsound = γ R T /MWmix (3.10)

Regarding the transport properties, the computation of the mixture viscosity is

27
3. ESPSS: European Space Propulsion System Simulation

computed as follows [64]:

nchem
X y i · µi
µmix = (3.11)
nchem
i=1
X
yi + yj · φij
j=1
j6=i

where φij is the interaction parameter estimated with the following formulation:
"  0.5  0.25 #2  0.5
1 µi MWi 2MWj
φij = 1+
4 µj MWj MWi + MWj

Similarly to viscosity, the thermal conductivity for mixtures is computed as


follows:
nchem
X y i · λi
λmix = (3.12)
nchem
i=1
X
yi + yj · ψij
j=1
j6=i

The interaction parameter for the thermal conductivity is based on the one
computed for viscosity. The expression is as follows:
 
2.41 · (MWi − MWj ) · (MWi − 0.142 · MWj )
ψij = φij · 1 +
(MWi + MWj )2

3.1.6. Real Fluid - Perfect gas mixtures

The field of mixture of flows has been and is still today the subject of intensive
research. Various models exist in the literature that represent with variable
accuracy the chemico-physical phenomena that occur in mixed flows. There are
mainly two different mathematical formulations of mixed flows :
• The two-fluid models, where equations are written for mass, momentum
and energy balances for each fluid separately.

• The mixture models, where equations for the conservation of physical


properties are written for the two-phase mixture.
Mixture models have a reduced number of balance equations compared to two-
fluid models, and may hence be considered as simplifications in terms of math-

28
3. ESPSS: European Space Propulsion System Simulation

ematical and physical complexity. However, some mixture models, like the
Homogeneous Equilibrium Model (HEM) [27] that is used within this work, are
still of significant interest. The HEM formulation has indeed multiple advantages
with respect to the other models :

• its convective part is unconditionally hyperbolic, which is not the case for
other models;

• it is very similar to the Euler equations, thus it can benefit of all the numeri-
cal studies made for Euler equations;

• we do not need to derive nor to implement the mass, momentum and energy
transfer between phases, as they cancel each other out in the HEM mixture
formulation.

This is why this very simple formulation is still worth being used, despite its
inherent limitations.

The most important restriction is that the flow should be in equilibrium, or


at least close to it, in order for the HEM formulation to approximate correctly the
physical behaviour of the flow. By equilibrium one means here that both phases
have the same velocity, pressure and temperature.
The homogeneous equilibrium model of a mixture of two fluids is then ap-
plied. One fluid must be real or a simplified liquid, and the other a perfect
non-condensable gas (ncg). Assuming that:

• The Real Fluid in possibly subcooled liquid, saturated or superheated vapour


state and the ncg form a homogeneous mixture with a uniform temperature.

• The Real Fluid, if present, occupies the entire volume. ncg, if present,
occupies the same volume as the Real Fluid vapour according to the Gibbs
Dalton Law.

• If ncg and Real Fluid Liquid are present, the Real Fluid vapour is saturated
(relative humidity is equal to 1)

• If ncg is present, the Real Fluid liquid conditions are the subcooled condi-
tions corresponding to P and T. (Liquid phase pressure = Pvap + Pncg ),

29
3. ESPSS: European Space Propulsion System Simulation

• The ncg gas is insoluble in the Liquid Phase of the Real Fluid. There can be
no ncg if the volume is filled with the liquid phase of the Real Fluid

The following state equations are involved in presence of liquid:

ρliq , uliq = fstate (f luid, P, T )


where P = Pnc + Pvap (subcooled conditions)
ρvap , uvap = fsat (f luid, P, T )
where Pvap = fsat (f luid, T ) (saturated conditions)

“u” is the internal energy. Subscript “nc” denotes the non-condensable fluid. “f”
denote the corresponding pure fluid functions. In this system of equations, Pnc
and T are unknowns.
Assuming that the volume density, ρ, the non-condensable mass fraction, xnc ,
and the mixture energy, u, are known, the following closing equations allow the
calculation of the homogeneous temperature and the non-condensable pressure:

ρnc α = ρ xnc where ρnc = fstate (f luidnc , Pnc , T )


u = (1 − xnc )(x uvap + (1 − x)uliq ) + xnc unc

Applying the definitions of the void fraction (α = Vg / Vtot ) and quality (x =


Mvap / (Mvap + Mliq )), it is possible to find an expression for the quality appearing
in the equation above as a function of densities:

α = (ρliq − ρcond )/(ρliq − ρvap )


x = α ρvap /(ρliq − α(ρliq − ρvap ))

The variable ρcond = (1 − xnc )ρ refers to the condensable mass (liquid & vapour)
divided by the total volume. The non-condensable density ρnc and the vapour
density are referring to the gas volume and not to the total volume, so ρnc 6= xnc ρ.
For the sake of clarity, it can be seen that the last two equations are identities
introducing the definition of:

• Vg = Vvap = Vnc

30
3. ESPSS: European Space Propulsion System Simulation

• Vtot = Vg + Vliq

• α = Vg / Vtot

• x = Mvap / (Mvap + Mliq )

• ρ = (Mnc + Mliq + Mvap ) / Vtot

• ρcond = (Mvap + Mliq ) / Vtot

• ρnc = Mnc / Vg

• ρliq = Mliq / Vliq

• ρvap = Mvap / Vg

The heat capacity, viscosity and thermal conductivity are calculated as a mix-
ture of a liquid and a composed gas (the vapour and the non-condensable gas).
The mixture properties are calculated in a simple way (weighing the pure fluid
properties with the mass fractions):

Cp = xmix Cpgas + (1 − xmix )Cpliq (3.13)


µ = xmix µgas + (1 − xmix )µliq (3.14)
λ = xmix λgas + (1 − xmix )λliq (3.15)

where the mixture quality (xmix ) is defined as the mass ratio of gas (vapour + non
condensable)
xmix = α ρgas /(ρliq − α(ρliq − ρgas ))

The void fraction α has the same meaning than in a pure two-phase fluid, i.e.
the gas volume divided by the total fluid volume. The gas mixture properties are
calculated as follows:

ρgas = ρvap + ρnc


Cpgas = (ρvap Cpvap + ρnc Cpnc )/ρgas

31
3. ESPSS: European Space Propulsion System Simulation

Similarly, the gas mixture transport properties and sound speed are calculated
as follows:

µgas = (ρvap µvap + ρnc µnc )/ρgas


λgas = (ρvap λvap + ρnc λnc )/ρgas
2 2 2
vsound,gas = ρgas /(ρvap /vsound,vap + ρnc /vsound,nc )

All individual properties have been computed with pure fluid functions.
The sound speed is approximated as an equivalent two-phase mixture where
the vapour phase is in fact a mixture of a non-condensable fluid with 100% of
humidity:

2 2 2
1/vsound = (αρgas + (1 − α)ρliq )(α/ρgas /vsound,gas + (1 − α)ρliq /vsound,liq )

3.2. Fluid Flow 1D Library

fluid_flow_1d is an EcosimPro library for 1-D transient simulations of two-fluid,


two-phase systems. The most important features are the following:

• The conservation equations include gas, liquid and two-phase flow regimes
for ideal or real fluids. The working fluid(s) can be easily selected from a
large collection of fluids included in the fluid_properties library.

• The fluid phase will be automatically calculated. The homogeneous equilib-


rium model is used to calculate a real fluid under two phase conditions with
or without a non-condensable gas mixture. Absorption/desorption is not yet
considered.

• Flow inversion, inertia, gravity forces and high speed phenomena are con-
sidered in pipes, volumes and junctions, the pipes also incorporating an
area-varying non-uniform mesh 1-D spatial discretisation into n (input data)
volumes.

• Calculation of concentrated (valves) and distributed (pipes) load losses


including two-phase wall friction correlations.

32
3. ESPSS: European Space Propulsion System Simulation

• Heat transfer between the walls and the fluid. Multiple thermo-hydraulic
correlations and initialization options are included.

• Other special components such as check valves, pressure regulators, heat-


exchangers and Tees (for convergent and divergent flows) are available.

• 1-D Pipe flows can be simulated using some robust and accurate numerical
techniques upwind (Roe) or centred schemes.
Hydraulic or pneumatic systems where the heat transfer or system controls are
coupled will be easily evaluated with the fluid_flow_1d library. Cavitation and
priming phenomena under two-phase flow (with or without a non-condensable
gas travelling in a liquid) will be calculated in pipes or other components. Besides,
fluid_flow_1d will permit to analyse in great detail transient aspects due to
inertia (water-hammer) and bubble collapse (priming).

3.2.1. Components Classification

The components of the fluid_flow_1d Library are listed in Figure 3.1 below,
where the inheritance hierarchy is shown. In an ESPSS fluid network, every
component is either a resistive component or a capacitive component. A resistive
component receives the state variables (pressure, density, velocity, chemical com-
position and enthalpy) as input and gives back the flow variables (volumetric, mass
and enthalpy flows) as output. A capacitive component receives the flow variables
as input and gives back the state variables at output. To build a fluid network, the
user must connect resistive components to capacitive ones, alternatively. So, from
a computational point of view, components are divided into two classes:
• C (capacitive) elements, integrating the mass and the energy conservation
equations. The thermodynamic functions will be used to calculate the
complete thermodynamic state

• M (momentum) elements, calculating explicitly (inertia terms) the mass


flows between C elements. Reverse flow is allowed
This computational scheme prevents the appearance of algebraic loops and high
index DAE (Differential Algebraic Equations) in the mathematical model of the
pipe network.

33
3. ESPSS: European Space Propulsion System Simulation

Figure 3.1.: Components in the fluid_flow_1d library

The existing components of the fluid_flow_1d library have the following


types:

• Volumes and Heat-exchangers components are C elements.

• Junctions, Valves, Filter and Jun_TMD components are M elements.

• Bound components (VolPT_TMD, VolPx_TMD, etc) are also C elements.


(Here, TMD means time dependant). VolPsTsVs_TMD and VolPsTs_TMD
are M elements because they calculate mass flow. In TMD elements, state
variables are imposed in the experiment file (fixed or depending on the time)
or by means of control library components connected to its control ports.

• Pipe, Tube, HeatExchanger and Nozzle components are 1-D models with
dedicated numerical schemes comparable to a C element.

• Pipe_res and ColdThruster components are 1-D models with dedicated


numerical schemes comparable to an M element. The ColdThruster incor-
porates an internal valve component.

• Others topological components such as the Tee component are M elements


because they internally finish in junctions, even if they have some internal

34
3. ESPSS: European Space Propulsion System Simulation

C components.

The graphical symbols of the components provide information about the kind
of computational element to which each port is connected. Ports belonging to a C
element have a small dot in the middle of the arrow while ports belonging to an
M element are just represented by the arrow (see Figure 3.1). It is noted that the
Tube and Pipe components can simulate an area-varying non-uniform 1-D mesh,
as it is the case for the ColdThruster and Nozzle components, even though, for
simplicity, the symbol graphical representation does not indicate this capability
on Pipes and Tubes components.

3.2.2. Junction/Valve

This component represents a junction. It is a basic component where no mass


accumulation is considered. The mass and enthalpy flows of the inlet and outlet
ports are equal (no mass accumulation in junctions):

ṁ1 = ṁ2 = ṁ
ṁ1,nc = ṁ2,nc = ṁnc
ṁh1 = ṁh2 = ṁh

where indexes 1 and 2 refer to the connected fluid ports. For each port, “ṁ”,
“ṁnc ”, and “ṁh” are the total mass, non-condensable mass and the enthalpy flows
respectively. These flows are calculated here taking into account the flow direction,
so different temperatures and densities may exist at both sides of a junction.
The following momentum balance equation dynamically calculates the mass
flow per unit of area:
 
dG dA dG G|G|
(I1 +I2 ) A +G +lv = (P +0.5ρv 2 )1 −(P +0.5ρv 2 )2 −0.5(ζ+ζcrit )
dt dt dt ρup
(3.16)
where, P1 , P2 are the static pressures at port 1 and 2, calculated by the connected
volumes (0.5ρv 2 )1−2 represent the dynamic pressures at port 1 and 2, calculated
by the connected volumes. ρ and v are the mean density and speed at the
connected volumes; I1 , I2 represent half inertia of the connected pipe ends 1 and

35
3. ESPSS: European Space Propulsion System Simulation

2, respectively. A is the instantaneous valve cross section; lv = sqrt(Aref ), this


term (valve inertia) is very small but makes the equation no singular if A = 0.
G is the mass flow per unit of area, ζ is pressure drop coefficient and ρup is the
upstream gas/liquid mixture density.
The mass flow will be calculated as ṁ = GA. The use of G (mass flow per unit
or area) instead of ṁ (mass flow) allows a complete closing (A = 0) of the valve
without making the system of equations singular.
The pressure drop contribution is quadratic with mass flow: 0.5(ζ + ζcrit )G|G|/ρup .
This term would make the momentum equation singular at zero flow because
very small perturbations in pressure lead to non-negligible variations in mass flow
(∂G/∂P → ∞). Physically, what is happening is that pressures losses are linear
with mass flow for laminar regimes. To account for that, the quadratic term GkGk
is linearised for G < Glam in this way:

k(G)G G < Glam
G|G| =
G2 sign(G) G > Glam

Glam = µRelam A; k(G)G→0 → Glam

k(G) is a smoothing factor to assure continuous transitions between laminar


to turbulent regimes. µ is the upstream viscosity calculated by the connected
volumes, Relam is the minimum Reynolds number set at 2000 as default value.
The sonic flow limitation is taken into account by adding a correction factor
to the pressure loss coefficient, ζcrit , which limits the mass flow per unit area to
be less than or equal (≤) to the critical flow per unit area. First, the flow under
steady state conditions is obtained by cancelling the derivatives in the previous
momentum equation:
q
Gst = 2ρup [(P + 0.5ρv 2 )1 − (P + 0.5ρv 2 )2 ]

The added term is calculated in such a way that if the flow attempts to be
greater than critical flow, the following extra term will limit the flow to the critical
value:

36
3. ESPSS: European Space Propulsion System Simulation

ζcrit = max (Gst /Gcrit )2 − ζ, 0




where “Gcrit ” is the critical (sonic) flow per unit of area (ρc)crit calculated by
the capacity components (see section 3.2.3) connected by the junction.

3.2.3. Capacity/Volume

The Capacity component simulates a volume with several fluid ports named f[j].
It’s the basic capacitive component containing the mass and energy conservation
equations for this type of components.
Here below the general equations for a non-adiabatic variable volume. It is
assumed that the mixture (non-condensable plus main fluid in liquid, gas or two
phase conditions) has an homogeneous temperature.

Mass conservation
dρ dV X
V +ρ = ṁj (3.17)
dt dt
j∈P orts

Non-condensable mass fraction xnc conservation

dxnc
 
nc dρ dV X
ρV +x V +ρ = ṁnc
j (3.18)
dt dt dt
j∈P orts

Energy conservation

dρ dV du X
V u+ρ u + ρV = (ṁh)j + Q̇ − P dV (3.19)
dt dt dt
j∈P orts

u = total specific energy = ust + v 2 /2 (3.20)

where ρ, xnc and u are the fluid mixture (including two phase flow) density, the
non-condensable mass fraction and the total energy respectively; ṁj , ṁnc
j and
ṁhj are the mass and enthalpy flows at port j calculated at the connected resistive
type components (see section 3.2.2). V is the volume, which can change with time.
Assuming that the volume V and its rate of change are known, previous
conservation equations enable to calculate the derivatives of the mixture density,

37
3. ESPSS: European Space Propulsion System Simulation

mixture energy and non-condensable mass fraction. These variables can be


integrated, so they are known at any time.
Assuming thermodynamic equilibrium, the conservation equations are always
valid even if the fluid conditions are liquid, vapour or homogeneous two phase
flow. Then, the complete thermodynamic state (partial pressures, temperature,
quality ...) can be calculated using the pure fluid thermodynamic routines or the
homogeneous equilibrium model for mixtures of a non-condensable gas plus a
real fluid: FL_state_vs_ru
Volume average velocities are required for the total energy conservation equa-
tion, the evaluation of the wall frictional forces and of the wall heat transfer. For
the calculation of the average velocity, volumes are considered to have two sides,
side 1 and side 2. The total mass flow rate entering the volume at side 1 and 2 is:
X X
ṁ1,in = ṁj,in ; ṁ2,in = ṁj,in
∀ports in side1 ∀ports in side2

Port mass flows can be positive or negative. It is defined as positive when


entering the volume. The average velocity in the volume is defined as:

ṁin,1 − ṁin,2
v=
2ρA

where ρ is the average density in the volume, and A is the cross area of the
volume. This average volume velocity is transmitted to the ports. The effective
port velocity will be multiplied by the cosine of the port angle α because the
lateral velocities do not compute in the total pressure:

v(j) = v cos(αj )

The term Q̇ appearing in the energy conservation equation of the capacity


permits the exchange of heat through a thermal port. The walls (that can be
represented by thermal components) are not included in this component:

Q̇ = hf ilm Awall (tp.T − T )

where tp is the name of the thermal port (with one node) connected to the

38
3. ESPSS: European Space Propulsion System Simulation

Volume. “tp.T” behaves as the internal wall temperature, to be determined in the


connected thermal component. The film coefficient is calculated using empirical
correlations.

3.2.4. Tubes/Pipes

These components simulate area-varying non-uniform mesh high resolved 1D


fluid veins that exchanges heat with a 1D thermal port. They incorporate the
1D mass, energy and momentum equations in transient regime. The number of
volumes in which the pipe is discretised will be a parameter.
All kind of flows (compressible or nearly incompressible flows, single component
or two-component flows, single phase or two-phase flows) can be simulated by us-
ing the following system of governing equations, here in area-scaled conservation
form [138]:

Mass conservation:

∂ρ ∂ρvA ∂P
A + = −ρAkwall (3.21)
∂t ∂x ∂t

Non-condensable mass fraction xnc conservation:

∂ρxnc ∂ρvxnc A ∂P
A + = −ρxnc Akwall (3.22)
∂t ∂x ∂t

Momentum conservation:

∂ρv ∂[(ρv 2 + P )A]


 
1 dξ dA
A + =− ρ v|v|A + ρgA + P (3.23)
∂t ∂x 2 dx dx

Energy conservation:
!
∂ρE ∂ρvHA dQ̇w
A + = + ρgvA (3.24)
∂t ∂x dx

where ρ, xnc , P , u are the gas/liquid mixture density, the non-condensable mass
fraction, the pressure and the total energy respectively. A is the variable flow
area and v the velocity. This system of 4 equations represents the general case

39
3. ESPSS: European Space Propulsion System Simulation

of a mixture of two fluid components, for which the first one can be either one
phase or two-phase, and the second one is always a non-condensable gas. This
set of equations is closed by a thermodynamic equation of state (EoS), which is
described in the fluid properties library, and hereafter written under general form:

p = p(ρ, u) (3.25)

The choice of density ρ and internal energy u as independent thermodynamic


variables is the most efficient one regarding CPU-time when the EoS is left under
arbitrary form.
The different source terms are the following:

• In the first equation governing the mixture mass conservation, a source term
responsible for the wall compressibility effect of the mixture, determinant
in water hammer simulations, is included; kwall is the wall compressibil-
ity. Assuming linear elasticity for the pipe wall material, we have three
configurations:
pipe anchored with expansion joints throughout: kwall = Din /t/ME
 
Din 5
pipe anchored at its upstream end only: kwall = −η
t ME 4
Din
pipe fully anchored: kwall = (1 − η 2 )
t ME
ME is the Young’s modulus of elasticity, η is the Poisson’s ratio and t is the
wall thickness.
The wall compressibility shall be multiplied by ∂P/∂t to account for the
volume change. For this purpose is calculated from the current state variables
(density and energy) and the thermodynamic derivatives:
    
∂P ∂ρ ∂ρ ∂u P ∂ρ ∂ρ 1 ∂ρ
= − + 2 / −
∂t ∂t ∂h P ∂t ρ ∂t ∂P h ρ ∂h P

• In the second governing equation, the non-condensable mass conservation


(if any), a similar source term is included;

• In the third governing equation, the mixture momentum conservation, a


source term represents the friction (dξ , proportional to dx, is the pressure

40
3. ESPSS: European Space Propulsion System Simulation

drop coefficient derived by empirical correlations), another one takes into


account the gravity, and the last one is responsible for the area variation.
The equivalent distributed friction, ∆ξi is calculated as follows:
X ∆xi
∆ξi = kadd + hdcbend (αbend,j , Rbend,j , Di , ) + hdcfric (Di , , Rei )
Di
bend,j

where kadd is an input data representing concentrated load losses to be


distributed along the pipe; Function hdcbend calculates the bend pressure
drop coefficient; Function hdcf ric calculates the friction factor including
laminar and turbulent regimes.
“g” represents the gravitational acceleration, if any. It is computed as the
scalar product of the gravity vector (gx , gy , gz ) with the direction of the pipe
in the global axis system (∆x, ∆y, ∆z), which are the difference of position
of the tube tips.

• In the last governing equation, the mixture energy conservation, a source


term Q̇w takes into account the heat transfer with the wall when it is
included:

Q̇w = hf ilm dx[Pinner (tp_in.T − T ) + Pouter (tp_out.T − T )]

where Pinner and Pouter are the wet perimeters; tp_in, tp_out are the names
of the thermal ports connected to the tube with the same number of nodes
as the fluid vein.
The port temperatures tp_... T behave as the wall internal temperatures, to
be determined in the connected thermal components. The film coefficient
for each node is calculated using empirical two-phase correlations:

hf ilm = htc(x, Dh , λ, Re, P r . . .)


G Dh
Re =
µ
µ Cp
Pr =
λ

41
3. ESPSS: European Space Propulsion System Simulation

where G = ρv is the mass flow rate per unit area, Dh is the hydraulic
diameter of the pipe Dh = 4A/Pw . In case of circular cross section Dh is
equal to the geometric diameter of the cross section; otherwise it represents
a reasonable characteristic length of the cross section.
Another source term, ρgvA, takes into account the gravity work.

The tube and pipe components are discretised by either a centred or an upwind
numerical scheme. Figure 3.2 describes the pipe discretisation. The inner fluxes are
computed using one or the other of these schemes, and the first and last junctions
(1 and n + 1) ones are given by the fluid ports, as they are calculated at resistive
type components using momentum equation with sonic flow limitation. Note
that the first and last half-nodal inertia are included in the junction component
equations.

Figure 3.2.: Pipe discretisation

Using the centred scheme, a staggered mesh approach is applied, for which the
state variables (pressure, density, velocity, chemical composition and enthalpy) are
associated with the n nodes, and the flow variables (volumetric, mass and enthalpy
flows) are calculated at the internal junctions (each junction has associated two
half volume inertias). With this scheme, the various fluxes to be computed at the
inner junctions are simply the flow variables, except for the mixture momentum

42
3. ESPSS: European Space Propulsion System Simulation

flux that is associated to the n nodes:

fmass,j = ṁj
fnc,j = ṁnc
j (3.26)
fmom,j = (Pi + ρi vi2 + qni )Ai
fene,j = ṁHj

The momentum flux term includes an artificial dissipation term qni calculated as
follows:
ṁi+1 − ṁi
qni = −Damp vsound,i
A
As an alternative to the centred scheme, an adequate upwind scheme has been
developed to improve the discontinuities resolution of these transient two-phase
flows in quasi-1D pipe networks. Using that scheme, a collocated mesh approach
is applied, for which all variables are discretised on the n nodes, even the mass
flow. All the fluxes f (u) are discretised at the junctions and include a central
part and an upwind part, following the initial Roe scheme [123]. Please refer to
Appendix A for a detailed description of the scheme.

3.3. Turbomachinery Library

turbo_machinery is an EcosimPro library for the simulation of pumps, turbines


and compressors. The most important features are the following:

• Pump model provided with user-defined dimensionless turbo-pump charac-


teristic curves adapted to positive and negative speeds and flow zones

• Turbine and Compressor components provided with user-defined dimension-


less performance maps as a function of the reduced axial speed and pressure
ratio

• Special turbomachinery components allowing simple calculation of general-


ized performance maps as a function of the nominal performances and other
significant design data

43
3. ESPSS: European Space Propulsion System Simulation

• Programming of the turbomachinery components is non-dependant on the


working fluid type: the properties of the selected working fluid (calculated
inside the corresponding thermodynamic function) will depend on the fluid,
but the non-dimensional parameters of the performance maps are equally
defined for all kind of fluids.

The generalized performances maps of the turbo_machinery library allows


robustly analysing the transients during the start up and shutting down processes,
where the reduced axial speed and flow are far away from the nominal values. The
turbo_machinery components can be connected to fluid_flow_1d components
with the aim of simulate a complete rocket engine cycle.
The components of the turbo_machinery library are listed below, and repre-
sented in Figure 3.3:

- Compressor

- Pump

- Turbine

- Compressor generic

- Pump generic

- Turbine generic

- Pump vacuum

Two different types of pumps/turbines are available: one “generic” model if the
off-design characteristics are unknown and one specific model which can only be
used with tables, for well defined turbo machinery.
All these components behave externally as resistive elements (see section 3.2.1).
Components ports are resistive because they calculate the mass flow. Nevertheless,
every model of a turbo-machinery component includes an internal capacitive
element receiving/giving the mechanical work.
Since in turbopump-fed engine cycles only pumps and turbines are present, only
these two class of components will be discussed in this section.

44
3. ESPSS: European Space Propulsion System Simulation

Figure 3.3.: Components in the turbo_machinery library

3.3.1. Pump & Generic Pump

This component simulates a pump for liquids. It is provided with fixed or user-
defined dimensionless turbo-pump characteristic curves adapted to positive and
negative speeds and flow zones, and valid for several types of pumps.
Usually it is rather difficult to find pump curves including the non normal zones,
so this is the reason to include in this component a set of curves [25] covering all
zones of the pump operation for 3 different specific speeds: Ns = 25 corresponding
to a pure centrifugal pump, Ns=147 corresponding to mixed pump, and Ns=261
corresponding to an axial pump. The specific speed is defined as:
p
rpm Q/ns
Ns =
(TDH/nst )0.75

where, ns is number of suctions, nst is the number of stages, Q [m3 /s] is volumetric
flow and TDH [m] is the actual total dynamic head of the pump.
The pump model makes use of performance maps for head and resistive torque.
The pump curves are introduced by means of fixed 1-D data tables defined as
functions of a dimensionless variable θ that preserves homologous relationships

45
3. ESPSS: European Space Propulsion System Simulation

in all zones of operation. θ parameter is defined as follows:

θ = π + arctan(ν/n) (3.27)

where ν and n are the reduced flow and speed parameters respectively:

Q ṁin /ρin ω
ν= = n= (3.28)
QR QR ωR
The dimensionless characteristics (head and torque) are defined as follows:

TDH / TDHR τ / τR
h= β= (3.29)
n2 + ν 2 n2 + ν 2
τ and TDH are the torque and the total dynamic head respectively. Sub index R
means “rated” (nominal) conditions. The nominal torque is calculated from the
other nominal parameters:

g ρup TDHR QR
τR = (3.30)
ηR ωR

This method eliminates most concerns of zero quantities producing singularities.


To simplify the comparison with generic map curves, these relations are normal-
ized using the head, torque, speed and volumetric flow at the point of maximum
pump efficiency.
In case there are no user defined curves, the ones already implemented in the
component at different specific speeds will be used and interpolated as function of
the actual Ns and θ:

h = h_vs_theta_Ns(Ns, θ) dimensionless pump TDH


β = β _vs_theta_Ns(Ns, θ) dimensionless pump torque

using the definition of h and β , the actual torque τ and the pressure rise TDH
(expressed as the total dynamic head in meters) will be calculated.
The mechanical balance allows the calculation of the axial speed dynamically:


Imech = τshaf t − τ (3.31)
dt

46
3. ESPSS: European Space Propulsion System Simulation

where ω is mechanical speed, Imech is the mechanical inertia and τ is the torque
calculated using the non-dimensional performances pump.
The enthalpy flow rise is a function of the absorbed power while the evaluation
of the mass flow rate is performed through an ODE.



 (ṁ h)out = τ · ω − (ṁ h)in

(3.32)
 dṁ
= P + 12 ρv 2 out − P + 12 ρv 2 in − gρin · TDH
  
I ·

dt
In case performance maps of a particular pump are known, a different approach
is used; the pump curves (head and torque) are introduced by means of input data
tables:
Independent variables:

- Mass flow coefficient: φ+ = ṁ/(ρin ω)

Dependent variables:

- Head rise coefficient: ψ + = ∆P/(ρin ω 2 )

- Reduced torque: C + = τ /(ρin ω 2 )

The head rise and needed torque coefficients are computed with 1-D tables,
as a function of mass flow coefficient only. Rotational speed is not taken into
account. The pump model described here has the disadvantage of being less
general than the Pump_gen component: the coefficients used (φ+ , ψ + and C + )
are not dimensionless. Hence, the characteristics differ for each pump, even
for geometrically similar pumps (with impellers having the same angles and
proportions).

3.3.2. Turbine & Generic Turbine

This component simulates a turbine for gases. It is provided with calculated (no
input) but adjustable dimensionless characteristic curves adapted to positive and
negative speeds and flow zones, and valid for several types of turbines. Adjustable
dimensionless pressure ratio curves are function of beta design parameter, reduced

47
3. ESPSS: European Space Propulsion System Simulation

speed and inlet Mach number. Efficiency curves are function of reduced speed
and pressure ratio parameters.
The value of the power W is obtained using GasTurbo_pow function. This
function basically estimates the efficiency as a function of the reduced speed and
pressure ratio parameters. It also needs the inlet mass flow to calculate the power.
The rest of the arguments are input data (ηnom , Nnom ,...) or calculated values in
the connected pipes components (bound pressures and enthalpies, etc.).
As for the pump, the mechanical balance allows the calculation of the axial
speed dynamically:

Imech = τshaf t − τ
dt
The mechanical work (τ · ω) extracted from the turbine is simulated as an
enthalpy flow:
(ṁ h)out = (ṁ h)in − τ · ω

The inlet mass flow equation is expressed dynamically in accordance with the
turbine pressure drop as follows:
   
dṁ 1 2 1 2
I· = P + ρv − P + ρv − (Πnom − 1) Pin · dp_rel
dt 2 in 2 out

dp_rel is obtained using GasTurbo_dpTurb function. This function basically


estimates the relative pressure drop of the turbine as a function of the beta design
parameter, the reduced speed and the inlet Mach number.
ṁin
The inlet Mach number Min is related to the inlet mass flow: Min =
ρin Ain Co
Co is the sound speed calculated at the outlet volume or at inlet according to the
vsound_outlet option.
When specific turbine performance maps are introduced, the Turbine_Gen
component cannot be used anymore and the Turbine component is then necessary
adopted. Turbine map curves (dimensionless torque and mass flow coefficients)
are input data tables depending on the dimensionless speed and pressure ratio
coefficients.
The performance maps (mass flow coefficient and specific torque) are introduced
by means of 2D input data tables:
Independent variables:

48
3. ESPSS: European Space Propulsion System Simulation

r·ω
- Speed coefficient: N =
Co
- Total pressure ratio: Π = P01 / P02

Dependent variables:
ṁmap · Co
- Mass flow coefficient: Q+ =
r2 P01
τ
- Specific torque: ST =
r ṁmap Co

Independent variables are easily calculated because they are obtained from
the dynamic rotational speed and from the boundary pressures. Then, the mass
flow and torque will be computed interpolating into the 2D input data tables
representing the turbine maps. Compared to the generic component model, only
the mass flow equation is different. The inlet mass flow equation is expressed
dynamically in accordance with a time delay (inertia terms):

dṁ
τ· = (ṁmap − ṁ) ; τ = I · r2 /Co
dt

3.4. Combustion Chambers Library

comb_chambers is an EcosimPro library for the simulation of rocket engines. The


most important features are the following:

• The properties of the combustion gases (transport and heat capacity) are
obtained from the CEA [55] coefficients (see Section 3.1.1) for an arbitrary
mixture or chemical’s reactants. The equilibrium molar fractions at of a
mixture of reactants are derived from the Minimum Gibbs energy method

• Non adiabatic 1D Combustor component: the equilibrium combustion gases


are calculated using previous capabilities. The chamber conditions will be
derived from the general transient conservation equations along a 1D spatial
discretisation

• Inclusion of another more advanced non-equilibrium, non adiabatic 1D


combustor component where a model for the liquid droplets evaporation
and for the global reaction time are also considered. The non-equilibrium

49
3. ESPSS: European Space Propulsion System Simulation

model is only a first approach; it does not include finite rate chemistry but
only time delay parameters

• Pre-burners and main combustion chambers are topologically built by means


of a combustor, two injectors and two cavity components. They can work
either under liquid, two-phase or gas injection conditions. The bubble
collapse calculation is included in the model of cavities

• The combustion gases generated in a chamber can be conducted (using


standard fluid_flow_1d components) to the turbines or even to other
chambers where any of the previously combusted gases will be considered a
new reactant.

• Cooling jacket component: Several models are available: one with a com-
plete 3D wall temperature distribution and another also including the injec-
tion tores

• Modelling of solid propellant starters, igniters and thermal coating protection


are available using combustor components

• Ideal or non adiabatic exhaust nozzles provided with a 1D spatial discreti-


sation. A special nozzle component allows simple simulation of the film
cooling injection

The comb_chambers components can be connected with fluid_flow_1d,


tanks or turbo_machinery components for the simulation of a complete rocket
engine cycle. Models where one or more chambers are present (staged engines)
can be evaluated. This library permits to analyse in great detail the transients
during the start-up and shut-down processes, where the valve sequences are
decisive.
The numeric method used for the resolution of the subsonic sections of a
combustor is based on the transient conservation equations. ESPSS combustor
models calculate the chamber pressure and the mixture ratio as a function of the
combustor geometry (design parameters) and the physical boundaries whereas in
CEA code those variables were imposed. Advantages of the ESPSS methods are:

50
3. ESPSS: European Space Propulsion System Simulation

• Transient phenomena (including pressure/temperature peaks at the start up


and shut down processes) are taken in to account

• The number of implicit equations is reduced, the state variables being


dynamic

• The wall heat exchange (non adiabatic terms) and the pressure drops are
taken into account

• Transient formulation allows to include vaporization / non-equilibrium


phenomena, as it done in the Combustor_rate component

The drawbacks of this method are:

• The characteristic time (integration time step) can be very small. Neverthe-
less, numerical instabilities are normally smoothed, the integration being
faster than using an implicit method

• The total pressure is not strictly conserved along the 1D combustor volumes.
Typical errors are between 0.5 and 1%, mainly produced near the throat
where the Mach number is close to 1

Concerning the supersonic sections of nozzles, a resolution method based on


the transient conservation equations would have numeric problems (passage from
subsonic to supersonic regime, shock waves if non-adapted conditions, etc), so
a 1D quasi-steady implicit method has been implemented for these components,
including non-isentropic effects under frozen or equilibrium conditions.
The components of the comb_chambers library are listed in Figure 3.4, where
the inheritance hierarchy is shown. The existing components have the following
types (see section 3.2.1):

• Cavities, Regenerative Circuit and Nozzle components are C (capacitive)


elements

• Injectors and Chemical inflator components are M (resistive) elements

• Propellant fluid ports of Preburners and CombustChamberNozzle compo-


nents are capacitive, and can be connected to any fluid_flow_1d library

51
3. ESPSS: European Space Propulsion System Simulation

resistive component. Thermal ports should be connected to a Cooling Jacket


component or other thermal component

Figure 3.4.: Components in the comb_chambers library

ABS_Combustor, Injector and Inj_Cavity components should only be used for


building Combustion Chambers and Preburners because they must be intercon-
nected to each other.
Preburner components include a combustor, two injectors and cavities compo-
nents, and an outlet resistive fluid port where the mass flow is calculated. This port
must be connected to a fluid_flow_1d capacitive component so the combustion
gases can be conducted to another combustor.
CombustChamberNozzle components include a combustor, two injectors and
cavities components, and a non-adiabatic 1D Nozzle. They have a special exit port
so a new nozzle extension component (with or without film cooling injection) can
be connected.

52
3. ESPSS: European Space Propulsion System Simulation

Th_mux, Th_demux are multiplexer/demultiplexer components allowing split-


ting or jointing a vectorized thermal port into several ones, so that different sized
Cooling Jacket components can be connected to a chamber.

3.4.1. Injector Cavity

Inherited from a Capacity (fluid_flow_1d library, see Section 3.2.1) this compo-
nent represents the combustion cavities upstream the injectors with thermal ports
allowing heat exchange.
The formulation of the conservation equations are the same as in Section 3.2.3
they are valid even if the fluid conditions are liquid, vapour or homogeneous two
phase flow.
This component is also charged of the calculation of the propellant molar
fractions from the injected fluids:

Nchem = 1; (case of pure fluid)


NX
chem
Nk = yk ; MWmix = yk MWk ; (Case of previously burned gases)
k=1

In case of a pure fluid, chem is the chemical corresponding to the main fluid. In
case of previously combusted gases, yk are the molar fractions at the inlet of the
cavity calculated by an upstream combustor. MWk is the molecular weight of the
chemical constituent k .

3.4.2. Combustor Equilibrium

In the combustion chamber the merged, mixed and atomized propellants are va-
porized and burned. In doing so, the chemical bound energy from the propellants
is transformed into thermal energy. Hence, it follows an increase of the combus-
tion chamber temperature Tc , which also involves a pressure increase Pc in the
chamber.
This component represents a non adiabatic 1D combustion process inside a
chamber for liquid or gas propellants. The transient conditions (pressures, tem-
peratures, mass flows and heat exchanged with the walls) will be derived from
general 1D transient conservation equations.

53
3. ESPSS: European Space Propulsion System Simulation

The equilibrium combustion gases are calculated using the minimum Gibbs
energy method as a function of the propellant’s mixture molar fractions and
enthalpies, and the chamber pressure.
A mixture equation between the injected propellants and the combustion gases
is applied. From the definition of the mixture ratio (MR) and derivation, the
following dynamic equation gives the MR evolution:

MR = Massox /Massf u 
 dMR (ρV )chamber
(ρV )chamber  ⇒ ṁox = MR · ṁf u + (3.33)
Massf u =  dt 1 + MR
1 + MR

The implementation of the 1D set of equations is done according to a staggered


grid in which the P /ρ/x variables are defined in the centre of the volumes and the
mass flows at the junction between the volumes.

Gas mixture mass conservation equation:

∂ρ ∂ρvA
A + =0 (3.34)
∂t ∂x

Gas mixture momentum conservation equations:

∂ρv ∂[(ρv 2 + P )A]


 
1 dξ dA
A + =− ρ v|v|A + P (3.35)
∂t ∂x 2 dx dx

Gas mixture energy conservation equation:


 
∂ρE ∂ρvHA dq̇w
A + = (3.36)
∂t ∂x dx

At injector level (i=0), the junction mass flows will be calculated by the Injector
components. The effective liquid mass flow entering into the chamber will be
computed as follows:

• Burning = FALSE, only the injected vapours and non-condensable gases


contribute to the chamber pressurisation. Injected liquids are supposed to be
lost

54
3. ESPSS: European Space Propulsion System Simulation

• Burning = TRUE, it is supposed that all the injected liquid will be vaporised
within a delay time, τv (injected gas is not modified):

dṁf u,liq Tn − Ttr,f u


   
= (1 − xf u )ṁf u tanh 10 − ṁf u /τv
dt Ttr,f u
dṁox,liq
   
Tn − Ttr,ox
= (1 − xox )ṁox tanh 10 − ṁox /τv
dt Ttr,ox

where ṁox,liq , ṁf u,liq are the effective injected mass flow; ṁox , ṁf u are
the total injectors’ mass flows; xf u and xox are the gas (vapours and non-
condensable gases) mass fractions in the cavities. The hyperbolic tangent
term is added to produce a relaxation of the injected liquid mass flow at
very low temperatures (ignition process)

Total enthalpies hjun are calculated using the upstream cell conditions: hjun,i =
hi−1 = (u + P/ρ)i−1 .
Again, at injector level the enthalpy flows will be computed using the effective
liquid and gas mass flows multiplied by the corresponding cavity enthalpy (liquid
and vapour). In this way injection conditions (liquid, gas or two-phase flow) will
be taken into account.
The starter terms (starter_m and starter_mh) will be added to the mass and
energy conservation equations of the first chamber volume. The composition of
the solid propellant gases is an input data within a predefined set of chemicals
starter_mh = f(starter_T, powder_composition);
starter_m, starter_T and powder_composition being input data.
In the momentum fluxes an artificial dissipation qn is added, and is calculated
as follows:
ṁjun (i + 1) − ṁjun (i)
qni = −Damp vsound (i)
A
Damp is a global input data of the fluid_flow_1d library. Momentum equa-
tions are applied to the exit of any volume in which the combustor is discretised
with the exception of the last volume that should end with the throat to avoid
numerical problems (transition from subsonic to supersonic flow). The outlet mass
flow will be calculated at the throat.
Node velocities v are calculated using the adjacent junction mass flows and the

55
3. ESPSS: European Space Propulsion System Simulation

mean junction densities:


 
1 ṁjun,i−1 ṁjun,i ṁjun,1
vi = + i = 2, n; v1 = (3.37)
Ai ρi + ρi−1 ρi + ρi+1 A1 ρ1

For the first volume, it is supposed that only the outlet junction will account for
the node velocity. For the last volume, the throat density is used.
The vapours and the non-condensable conservation equations take into account
the mixture process:

∂xf u ρ ∂xf u ṁjun


 
A+ =0
∂t ∂x
 
∂xox ρ ∂xox ṁjun
A+ =0
∂t ∂x
∂xnc ṁjun
 nc 
∂x ρ
A+ =0
∂t ∂x
 
∂xpw ρ ∂xpw ṁjun
A+ =0
∂t ∂x

where, xf u is the reducer vapour mass fraction, and xox is the oxidizer vapour
mass fraction ; xnc is the non-condensable mass fraction and xpw represents the
solid propellant gases mass fraction.
The mass flow of the injected vapours (mf u , mox ) and injected non-condensable
gases will be added as source terms to the respective conservation equation of
the first chamber volume. In the same way, the mass flow of the solid propellant
gases, starter_m, will be added as source a term to the solid propellant gases
conservation equation of the first chamber volume.
Under burning conditions, only the mass fractions of the first chamber volume
will be used to compute the molar fraction of the reactants. Subsequent volumes
will use as reactant the product of the upwind volume.
For each chamber volume the combustion gases properties (product’s molar
fraction, heat and transport properties) are calculated using the Minimum Gibbs
energy method as a function of the propellant molar fractions, pressure and
specific enthalpy:

56
3. ESPSS: European Space Propulsion System Simulation

Reducer contribution:
NX
chem
yk,f u
Nk,i = xf u,i ; MWmix,f u = yk,f u MWk,f u
MWmix,f u
k=1

Oxidiser contribution:
NX
chem
yk,ox
Nk,i = Nk,i + xox,i ; MWmix,ox = yk,ox MWk,ox
MWmix,ox
k=1

Solid propellant gases contribution:

NX
chem
yk,pw
Nk,1 = Nk,1 + xpw,i ; MWmix,pw = yk,pw MWk,pw
MWmix,pw
k=1

Non-condensable gases contribution:

xnc
i
Nnc,i = Nnc,i +
MWnc

N chem is extended to any chemical treated by the fluid_properties library;


MWk is the molecular weight of the chemical constituent k; xnc
1 is the mass
fraction of non-condensable gas at volume 1.
xf u,i , xox,i , xpw,i are mass fractions of vapours and solid propellant gases;
yk,f u ; yk,ox ; yk,pw are the molar fraction of chemical k of the reducer (oxidizer)
mixture and Nk,1 is number of moles of the chemical constituent k of the reactant
mixture at volume 1.
The propellants (reducer and oxidizer mixtures) molar fractions have been
calculated by the cavity components. Propellants can be formed by any allowed
fluid mixture using the fluid_flow_1d library, including those of a previous
combustion.
Then, the number of moles, Nk is normalized. With this entry calculated and
using the “dynamic” enthalpy value obtained from the conservation equations of
the first combustor volume it is possible to call the Minimum Gibbs energy method
to obtain the equilibrium temperature and the molar fraction of the products:

(yk_eq , Teq ) = fminGibbs (Nk,1 , h1 − v12 /2, P1 )

57
3. ESPSS: European Space Propulsion System Simulation

Two possibilities are foreseen calling previous function:

- Equilibrium

- Frozen flow

In the last case (no ignition), molar fractions remain constant: yk_eq,1 = Nk,1 .
Under frozen conditions, the molar fractions of the subsequent volumes will be
calculated as for the first volume. The temperature calculation from the enthalpy
value will require an iteration procedure, in this case lesser complicated than in
equilibrium conditions.
Under equilibrium conditions, subsequent volumes will consider that the molar
fraction of the products of the previous volume will act as the inlet propellant
mixture, so the Minimum Gibbs energy method can be applied to any combustor
volume.
(yk_eq , Teq )i = fminGibbs (Nk,i−1 , hi − vi2 /2, Pi )

The effective combustion gas constants (Ri , Cpi , λi , µi ) will be derived using
the mixture properties equations as a function of yk_eq,1 , see Section 3.1.5. The
pressure is obtained from the perfect gas state equation:

Pi = ρi · Ri · Ti · η (3.38)

where η is the combustor efficiency. Note that the pressure equation and the
Minimum Gibbs function become an algebraic loop.
Finally, the molar fraction of the products of the last volume will be transmitted
to the outlet port to be used by the Nozzle component or in another possible
chamber.
The term q̇w appearing in the energy conservation equation permits the ex-
change of convective and radiative heat through a thermal port. The walls (that
can be represented by thermal components or by the Cooling Jacket component)
are not included in this component:

q̇w = hc Awet (Taw − tp.T ) + σAwet (T 4 − tp.T 4 ) (3.39)

tp is the name of the thermal port (with n nodes in axial direction) connected to

58
3. ESPSS: European Space Propulsion System Simulation

the Combustor. σ is the Stefan-Boltzmann constant = 5.67.10-8 [W ∆m−2 ∆K −4 ].


The heat exchanged with the fluid is transmitted through this port: tp.q̇(i) = q̇w,i ;
tp.T(i) behaves as the internal wall temperature, to be determined in the connected
wall component. Taw is the adiabatic wall temperature defined as:
   
0.33 γ − 1 2 Cpλ
Taw = T 1 + Prref M ; Prref =
2 µ ref

The reference conditions (ref) are calculated at a temperature halfway between


the wall and the free stream static temperature. It is supposed that the mixture
composition do not change between these two temperatures. The film coefficient
for each volume is calculated using empirical correlations according to Bartz [9]:
0.6 0.1
λref
 
πDth
hc = 0.026 µ0.2
ref Cp0.4 0.8
ref (ṁth ) /A
0.9
(3.40)
µref 4Rcurv

where, µref : viscosity of combusted gases at volume no. i and Tref temperature
λref : conductivity of combusted gases at volume i and Tref temperature
Rcurv : Curvature radius of the throat
Dth : Throat diameter
ṁth : Throat mass flow
A: Cross section at volume i.

3.4.3. Combustor rate

This component represents non-equilibrium, non adiabatic quasi 1-D combus-


tor component for liquid or gas propellants. The transient chamber conditions
(pressures, temperatures, mass flows and heat exchanged with the walls) will
be derived from general quasi 1-D transient conservation equations. The non-
equilibrium model is only a first approach; it does not include finite rate chemistry
but only time-delay parameters.
The mass, energy and momentum equations include those of the equilibrium
Combustor component plus the ones concerning the vaporization model. Liquid
phase conservation equations are also included.

59
3. ESPSS: European Space Propulsion System Simulation

Gas mixture mass conservation equation:

∂ρ ∂ρvA ṁvap_f u Af u ṁvap_ox Aox


A + = + (3.41)
∂t ∂x Vf u Vox

Gas mixture momentum conservation equations:

∂ρv ∂[(ρv 2 + P )A]


 
1 dξ dA
A + =− ρ v|v|A + P (3.42)
∂t ∂x 2 dx dx

Gas mixture energy conservation equation:

∂ρE ∂ρvHA dq̇w q̇vap_f u Af u q̇vap_ox Aox


A + = + + (3.43)
∂t ∂x dx Vf u Vox

where, ρ and E are the gas mixture density and total energy; v is the mean
velocity; ṁvap_ox and mvap_f u represent the vaporized liquid mass flow of oxidizer
and of the reducer.

At injector level, the gas mass and enthalpy flows (mjun,0 ) will be calculated
by the Injector components taking into account the quality calculated in the
Cavities components. The liquid contributions will be considered in the droplets
conservation equations.

ṁf u = ṁf u,inj xf u ; ṁox = ṁox,inj xox

where ṁox,inj , ṁf u,inj are the injectors’ mass flow (liquid or gas) and xf u and
xox are the vapour mass fractions in the cavities. Then ṁjun,0 = ṁf u + ṁox . An
hyperbolic tangent term is added to produce a relaxation of the injected vapour
mass flow at very low temperatures.
Since a centred scheme with a staggered mesh is adopted, total enthalpies hjun
are calculated using the upstream cell conditions: hjun,i = hi−1 = (u + P/ρ)i−1 .
Starter terms are in included in the same way as in the combustor_eq compo-
nent. Also, the node velocities v are calculated as in the combustor_eq component.
The burned gases production rate does not contribute explicitly to the gas
mixture equation because the mass is conserved in the chemical reactions (no
condensation).

60
3. ESPSS: European Space Propulsion System Simulation

Vaporization flows, mvap , and enthalpies flows, q̇vap (source term for the gas
mixture conservation equations), are calculated by the droplet vaporization models.
Two models are available:

User defined model :


The vapour mass flows in each volume are calculated assuming a characteristic
vaporization time modulated with user defined vaporization factors:

ṁvap_f u, = fvap Mliq_f u /τvap ; ṁvap_ox = fvap Mliq_ox /τvap


q̇vap_f u = ṁvap_f u h(Tliq_f u ); q̇vap_ox = ṁvap_ox h(Tliq_ox )

where, Mliq is the liquid mass at the volume i; Tliq are the liquid droplets tem-
peratures, τvap is the characteristic vaporization time and fvap represents the
vaporization factor (time dependant input data) at the volume i.

Droplets vaporization model :


Assuming a very thin saturated layer between the droplets and the surrounding
gases, the conservation equations establish that the sum of convective heat plus
enthalpy mass flow are the same at both sides of the layer.
Then, the following set of equations is applied at each combustor volume i,
allowing the calculation of the mass and energy exchanges through this layer:
  

 hc (T − Tsat_f u ) + hc,liq_f u (Tliq_f u − Tsat_f u )
ṁ = A

vap_ f u liq _f u

(hvap_f u − hliq_f u )










q̇
vap_f u = ṁvap_f u hvap_f u − Aliq _f u−gas hc (T − Tsat_f u )





hc,liq_f u = 2 λliq_f u /Ddroplet_f u











Aliq_f u = fvap 6 Mliq_f u /ρliq_f u /Ddroplet_f u

61
3. ESPSS: European Space Propulsion System Simulation
  

 hc (T − Tsat_ox ) + hc,liq_ox (Tliq_ox − Tsat_ox )
ṁ = A

 vap_ox liq _ox
(hvap_ox − hliq_ox )










q̇
vap_ox = ṁvap_ox hvap_ox − Aliq _ox−gas hc (T − Tsat_ox )





hc,liq_ox = 2 λliq_ox /Ddroplet_ox











Aliq_ox = fvap 6 Mliq_ox /ρliq_ox /Ddroplet_ox

where, T are the gas temperatures, hc is the heat exchange coefficient at gas side.
Same value as for the wall is used. Tsat is the saturation temperature calculated
at the partial vapour pressure of volume i; hvap and hliq represent the saturation
enthalpies calculated at the partial vapour pressure of volume i;
Ddroplet is the mean droplets diameter and Aliq is the equivalent exchange area
between the droplets and the gas.
We point out that the droplet diameter has been modulated by the vaporization
factor. This factor is an input data depending on the time and on the volume
number.
In theory, assuming a known droplet size at the injection plate, the droplet
diameter evolution could be determined by “simple” equations relating the evapo-
rated mass flow with the liquid mass conservation equations. Nevertheless, due
to the high penetration and break up of the liquid jets, it seems more realistic to
assume a known droplet size at each chamber volume, the number of droplets
being determined by the current liquid mass.
In both methods (used defined and droplet model), the vapour mass flow is also
weighted with a factor to prevent vaporisation at very low temperatures (frozen
liquid):  
T − Tsat
ṁvap = ṁvap tanh 10
Tsat
In normal situations, T > Tsat , making the value of the hyperbolic tangent be
equal to one.

Burning Rate :
The burned gases mass flow (second source term for the vapour conservation

62
3. ESPSS: European Space Propulsion System Simulation

equations, see Eq. 3.44a and Eq. 3.44b) is calculated assuming a global characteris-
tic burning time [9]. It is also supposed that any species (vapour or burned gas)
present in gas mixture contributes to the global reaction rate, so the burning rate
will be proportional to the total gas mixture density:

ρV
ṁbu = fbu
τbu

where τbu is the characteristic burning time and fbu is the burning factor at the
volume i.
The burning factors are automatically set to one if the burning conditions are
true: mixture ratio within the allowed limits and ignition flag activated. Otherwise
the burning factors are set to zero.

vapours / non-condensable / solid propellant gases mass equations :


The vapours and non-condensable mass conservation equations take into account
the burned gases production and the vaporization terms previously calculated:

∂(xf u ρ) ∂(xf u ρvA) ṁvap_f u Af u xf u Af u


+ = − ṁbu (3.44a)
∂t ∂x Vf u Vf u
∂(xox ρ) ∂(xox ρvA) ṁvap_ox Aox xox Aox
+ = − ṁbu (3.44b)
∂t ∂x Vox Vox
nc nc
∂(x ρ) ∂(x ρvA)
+ =0 (3.44c)
∂t ∂x
∂(xpw ρ) ∂(xpw ρvA)
+ =0 (3.44d)
∂t ∂x

The mass flow of the injected vapours (ṁf u , ṁox ) and injected non-condensable
gases will be added as source terms to the respective conservation equation of
the first chamber volume. In the same way, the mass flow of the solid propellant
gases, starter_m, will be added as source a term to the solid propellant gases
conservation equation of the first chamber volume.

63
3. ESPSS: European Space Propulsion System Simulation

Liquids conservation equations :


The liquid mass and enthalpy flows are calculated assuming that vgas = vliq and
neglecting Cpliq derivatives:

∂Mliq_f u ∂(Mliq_f u v)

+ = −ṁvap_f u


∂t ∂x








∂ ∂


(M T )liq_f u + [(M T )liq_f u v] = −qvap_f u /Cpliq_f u


 ∂t ∂x

∂Mliq_ox ∂(Mliq_ox v)




 + = −ṁvap_ox
∂t ∂x








 ∂ (M T )liq_ox + ∂ [(M T )liq_ox v] = −qvap_ox /Cpliq_ox



∂t ∂x

Combustion gases properties calculation :


It is supposed that any molar fraction follows a global reaction rate accordingly
with the previously mentioned burning time:

dyk,bu (yk,eq − yk,bu )


=
dt τbu

where yk,eq is the equilibrium molar fraction of the chemical constituent k at each
volume i. yk_bu is the actual burned molar fraction of the chemical constituent k
at each volume i, and τbu is characteristic burning time.
For each chamber volume i, the combustion gases equilibrium composition
(needed for the calculation of the actual burned gases composition) is calculated
using the Minimum Gibbs energy method as a function of the gas mixture molar
fractions, pressure and the enthalpy.
The gas mixture molar fractions in each volume are calculated as follows:
Reducer vapours contribution:

NX
chem
yk,f u
Nk = xf u ; MWmix,f u = yk,f u MWk,f u
MWmix,f u
k=1

64
3. ESPSS: European Space Propulsion System Simulation

Oxidiser vapours contribution:

NX
chem
yk,ox
Nk = Nk + xox ; MWmix,ox = yk,ox MWk,ox
MWmix,ox
k=1

Solid propellant gases contribution:

NX
chem
yk,pw
Nk = Nk + xpw ; MWmix,pw = yk,pw MWk,pw
MWmix,pw
k=1

Non-condensable gases contribution:

xnc
Nnc = Nnc +
MWnc

Burned gases contribution:

NX
chem
yk,bu
Nk = Nk + xbu ; MWmix,bu = yk,bu MWk,bu
MWmix,bu
k=1

The vapour (reducer and oxidizer mixtures) molar fractions have been calculated
by the cavity components and can include any allowed fluid mixture using the
fluid_flow_1d library, including that of a previous combustion. The burned
molar fractions, yk_bu , are calculated dynamically.
Once the number of moles of the reducer/oxidizer/burned gases mixture has
been evaluated, and using the “dynamic” enthalpy value obtained from the con-
servation equations, it is possible to call to the Minimum Gibbs energy method to
obtain the equilibrium combustion gases composition:

(yk_eq , Teq )i = fminGibbs (Nk,i , hi − vi2 /2, Pi )

Two possibilities are foreseen calling the previous function: Equilibrium and
frozen flow. In the last case (no ignition) the molar fractions remain constant:
yk_eq,i = (Nk,i ).
The effective combustion gas constants (Ri , Cpi , λi , µi ) will be derived using
the mixture properties equations as a function of yk_bu,i see Section 3.1.5.
The pressure is obtained from the perfect gas equation: Pi = ρi · Ri · Ti . With

65
3. ESPSS: European Space Propulsion System Simulation

respect to the Combustor_eq component, the pressure equation and the Minimum
Gibbs function do not become an algebraic loop because the actual molar fractions
(yk_bu,i ) of the mixture are dynamic variables.
The molar fraction of the products of the last volume will be transmitted to the
outlet port to be used by the Nozzle component or in another possible chamber.

3.4.4. Nozzle

This component represents 1D supersonic nozzle in quasi-steady conditions. Two


possibilities are foreseen in the main body of this component:

• an “ideal” nozzle using a variable gamma approximation

• a non-adiabatic, non-isentropic nozzle

The mathematical model explained below can be applied either to a complete


nozzle or to a nozzle extension. The important point is to know the inlet total
conditions calculated from the upstream enthalpy/entropy conditions and from
the mass flow calculated in the complete nozzle component:

- If the nozzle is connected to a chamber (case of a complete nozzle), then the


total conditions will be those of the exit of the chamber transmitted by the
nozzle port.

- If the nozzle is connected to another nozzle (case of a nozzle extension),


then the total conditions will be those of the exit of the upwind nozzle.
Inlet/outlet mass flows are the same.

The throat calculation (choked mass flow, see below) will be only implemented
in the complete nozzle.
The choked throat conditions (Pth , Tth and vth ) can be calculated with the
following three equations assuming that the total enthalpy and the static entropy
are known (those of the last combustor volume):

2
hch,tot = h(Nk,th , Tth ) + vth /2
sch = s(Nk,th , Pth , Tth )
p
vth = γth Rth Tth ; ṁth = Ath vth Pth /(Rth Tth )

66
3. ESPSS: European Space Propulsion System Simulation

where, γth is the isentropic coefficient. This is a function of Pth , Tth , Nk,th ; s is the
entropy, a function of Pth , Tth and Nk,th (see section 3.1.1); h is the static enthalpy.
This one is a function of Tth and Nk,th (see section 3.1.1).
Nk,th is the number of moles of the chemical constituent k of the burned gases
at throat. It is calculated from the Minimum Gibbs energy method:

(Nk , T )th = fminGibbs (Nk,ch , sch , Pth )

where sub index “ch” denotes the conditions at the exit of the combustor. Two
possibilities are foreseen calling the previous function: Equilibrium and frozen
flow. In the last case, molar fractions remain constant. The equations above are
solved iterating in pressure and in temperature.
To cover subsonic conditions, the effective throat mass flow is calculated using
a dynamic momentum equation (see Section 3.2.2) where the left pressure cor-
responds to the combustor exit, and the right one to the external pressure. Of
course, under normal steady conditions, the dynamic mass flow will be limited to
the critical mass flow, ṁth , previously calculated.

“Ideal” supersonic nozzle :


The nozzle is divided in sections. Assuming an isentropic frozen expansion
between the node i and i+1 the temperature and pressure for each section can be
calculated as follows:
 
Ttot γi − 1 Ptot γi
θi = =1+ Mi2 ; δi = = θ γi −1 (3.45)
Ti 2 Pi

where, γi is the burned gases isentropic coefficient at section no.i. This is a


function of T
Mi is burned gases Mach number at section no.i.
Ttot is the burned gases total temperature (input).
Ptot is the burned gases total pressure (input).
The closing condition to calculate the Mach number knowing the area ratio is:

  2(γγi +1
Ai 2θi i −1)
Mi = ; γi = f (Ti ) (3.46)
Ath γi + 1

67
3. ESPSS: European Space Propulsion System Simulation

The equations above are solved iterating in Mach and in temperature.

Non-adiabatic, non-isentropic supersonic nozzle :


It is assumed that the non-adiabatic process takes place into two separated steps:

Heat losses :
From section i to i+1, the following losses in enthalpy and entropy will take place:

htot,i+1 = htot,i − q̇w,i /ṁth (3.47)


si+1 = si − q̇w,i /ṁth /Ti (3.48)

where q̇w,i is calculated with Equation 3.39 (using Bartz correlations). For the first
nozzle station (i = 1) the total enthalpy and the static entropy are known (those
of the exit of upstream component).

Expansion :
Assuming now that the isentropic relations are valid we have:

2
htot,i+1 = h(Nk,i+1 , Ti+1 ) + vi+1 /2 (3.49)
si+1 = s(Nk,i+1 , Pi+1 , Ti+1 ) (3.50)
ṁth = ρi+1 vi+1 Ai+1 (3.51)

Nk,i is the number of moles of the chemical constituent k at the nozzle sta-
tion i. It is calculated from the Minimum Gibbs energy method: (Nk , T )i+1 =
fminGibbs (Nk,i , si , Pi+1 ).
The three equations above with three unknowns (pressure, temperature and
speed) are solved iterating in pressure and temperature. Two possibilities are
foreseen calling the previous function: Equilibrium and frozen flow. In the last
case, molar fractions remain constant.
The following numeric derivatives are done to calculate the isentropic coefficient

68
3. ESPSS: European Space Propulsion System Simulation

at equilibrium conditions:

∂ ln ν
= 1 + ln Ntot,dT / ln(1 + )
∂ ln T
∂ ln ν
= 1 + ln Ntot,dP / ln(1 + )
∂ ln P

where Ntot,dT , Ntot,dP are the total number of moles (increments with respect to
one) due to separated perturbations in P/T (∆P = P ; ∆T = T,  = 1e − 4):


MinGibbsEnergy_PT mix, Nk , T (1 + ), P, Nk,dT , Ntot,dT

MinGibbsEnergy_PT mix, Nk , T, P (1 + ), Nk,dP , Ntot,dP

Nk is the number of moles of the chemical constituent k at equilibrium con-


ditions. Nk,dT , Nk,dP are the number of moles after perturbation in T, P. The
calculation of the specific heat at constant pressure, Cp, has two terms: Cpf r and
Cpre . The frozen one is calculated as in Section 3.1.1 and the reactive term is
calculated as follows:

Cp = Cpf r + Cpre
X  ∂ ln Nk  hk
Cpre = · Nh
∂ ln T T MWmix
k

where, 
ln(Nk,dT /MWmix,dT ) − ln(Nk /MWmix )
 
∂ ln Nk
=
∂ ln T ln(1 + )
Then calculation of the specific heat at constant volume and the isentropic
coefficient is:

R (∂ ln ν/∂ ln T )2
Cv = Cp + ·
MWmix ∂ ln ν/∂ ln P
Cp
γ=
Cv (∂ ln ν/∂ ln P )

Under frozen conditions, (∂ ln ν/∂ ln P ) is equal to one.


Regarding the heat exchanged with the walls, it is used the same formulation
present in the Combustor model (see Equation 3.39). In the case of an ideal nozzle
it is supposed that the heat exchanged with the walls is a small quantity with

69
3. ESPSS: European Space Propulsion System Simulation

respect to the enthalpy flow, so the isentropic equations are supposed still valid.
In order to calculate the thrust of the engine, here below the expressions needed
to calculate the thrust F and the Isp at section i:
p
Fi = ṁth Mi γi Rgas Ti + (Pi − Pout ) Ai
ISPi = Fi /ṁth

where Pout is the external boundary condition in pressure; Rgas represents the
burned gases constant at nozzle exit ṁ is the mass flow at throat.

3.4.5. Cooling Jacket components

These components represent a Regenerative Circuit of a Chamber. Two models


are here described: the CoolingJacket component and the CoolingJacket_simple
component. For the first one a 3-D geometry (built by means of several 3-D walls
around the channels) is taken into account. For the second one, a simplified wall
geometry is considered.

Cooling Jacket component

It is constructed by aggregation of one Tube (fluid_flow_1d library) representing


the channels and five 3-D walls around them. The cooling jacket is divided into a
variable number of sections in axial direction. Every section is made of:

• one fluid node of the Tube component (fluid_flow_1d library), which is


simulating the cooling channels

• five slides of the wall_3D components, which are simulating the metallic
walls. They are arranged according to Figure 3.5

Channels: they are simulated by only one Tube component making its “num”
data equal to the number of channels. The mesh size of each node is function of
the axial position through the non-dimensional geometry tables.

70
3. ESPSS: European Space Propulsion System Simulation

Figure 3.5.: Cooling jacket wall mesh [43]

The rectangular channel geometry (widths, heights, wet areas) is similarly


calculated but using the interpolated widths and heights values as follows:

Awet,i = 2 · (ai + bi ) · lch,i


ai = wch · interp(xi /L, wc_vs_L)


b = t · interp(x /L, tc_vs_L)
i ch i

Heat conduction: the wall_3D components used to simulate the walls will
calculate the heat conduction in every direction including the axial direction.
This thermal component features thermal ports in radial and in azimuth di-
rections allowing an exact calculation of heat conduction through the channel
corners. The walls are divided in 5 different 3-D components as shown in Fig-
ure 3.5. Each component has a 3-dimensional discretisation in tangential, radial
and longitudinal direction (dx, dy , dz ), respectively.
The formulation for this component is the typical one for 3-dimensional con-
duction elements; the thermal capacitance for each volume is defined as:

Ci,j,k = ρ Cp(i,j,k) dx dy dz (3.52)

71
3. ESPSS: European Space Propulsion System Simulation

the internal heat flows are evaluated by:



q̇x(i,j,k) = ki,j,k dy dz (Ti−1,j,k − Ti,j,k )/dx


q̇y(i,j,k) = ki,j,k dx dz (Ti,j−1,k − Ti,j,k )/dy (3.53)


q̇
z(i,j,k) = ki,j,k dx dy (Ti,j,k−1 − Ti,j,k )/dz

while the energy equation is:

dTi,j,k
Ci,j,k = q̇x(i,j,k) − q̇x(i+1,j,k) + q̇y(i,j,k) − q̇y(i,j+1,k) + q̇z(i,j,k) − q̇z(i,j,k+1) (3.54)
dt

As shown in Figure 3.5 only half channel has been considered because of
symmetry reasons, with left and right sides adiabatic:

q̇out,right_r = 0


q̇out,int_l = 0 q̇out,int_right_r = 0


q̇
out,ext_l = 0 q̇out,ext_right_r = 0

The heat flux to the external side is calculated as the sum of all wall nodes
connected to the ambient:
 
N
X N
X
q̇amb (i) = 2 · nch  q̇out,ext (j) + q̇out,ext_right (j)
j=1 j=1

The temperatures of the external walls (for each section i) in contact with the
exterior are supposed to be the same:
(
Tw,amb (i) = Tout,ext (j)
Tw,amb (i) = Tout,ext_right (j)

The same procedure is applied for the heat fluxes related to the internal side,
the combustion chamber:
  
 N
X N
X

q̇in (i) = 2 · nch q̇in,int (j) + q̇in,int_right (j)

 

j=1 j=1



 Tw,in (i) = Tin,int (j)

Tw,in (i) = Tin,in_right (j)

72
3. ESPSS: European Space Propulsion System Simulation

The wet surfaces of the walls (wall_int, wall_right and wall_ext) in contact
with the channel coolant, are supposed to be at three different temperatures (one
for each side): 
Tch,in = Twall,int


Tch,lat = Twall,right


T =T
ch,out wall,ext

The corresponding heat flux (for each section i) between the channel and the walls
around it are calculated as follows:
N

 X



 q̇ch,in = 2 q̇wall,int (j)
j=1



N


 X
q̇ch,lat = 2 q̇wall,right (j)
j=1



N



 X
q̇ch,out = 2 q̇wall,ext (j)



j=1

The channel heat fluxes are also calculated by the Tube component using as
input the wall/fluid temperatures and two phase hydraulic correlations for the
film coefficient evaluation. Then, a set of implicit non linear equations is formed
which is solved by EcosimPro. This also applies for the combustion chamber side,
where the heat fluxes are calculated by the Combustor component connected to
the CoolingJacket component.
Concerning the external side (ambient), the heat flow shall be defined according
to the thermal component connected to this port.

Cooling Jacket simple component

This component is constructed by aggregation of one Tube (fluid_flow_1d


library) representing the channels and three 1D bars around them.
The philosophy is the same as in the CoolingJacket component with the dif-
ference that here the five 3D-walls are replaced by three 1D-bars with only one
thermal node per section. So, the cooling jacket is divided into a variable number
of sections in axial direction. Every section is made of:

• one fluid node of the Tube component (fluid_flow_1d library), which is

73
3. ESPSS: European Space Propulsion System Simulation

Figure 3.6.: Simplified Cooling Jacket wall disposition [43]

simulating the cooling channels

• three slides of the bar_1D components, which are simulating the metallic
walls. They are arranged according to Figure 3.6

Figure 3.6 shows the different temperatures and heat fluxes within the cooling
channel. There Tc is the temperature of the hot combustion gas, TW is the temper-
ature of either the combustion chamber - or nozzle wall, TR is the temperature of
the cooling rib, TCH is the temperature of the cooling fluid, TS is the temperature
of the surface and Tamb is the ambient temperature. For the model it is assumed
that the temperatures are constant over the separate volumes.
Heat always flows from regions with a higher potential to regions with a lower
potential. Starting with the heat flux Q̇CC,W from the hot combustion gas to the
chamber - or nozzle wall, the heat flux then splits up in a smaller part which goes
into the channel ribs Q̇W,R and the main part which goes into the cooling fluid
Q̇W,CH . From the ribs the main part flows into the cooling channel Q̇R,CH , but
also a small part flows to the surface Q̇R,S .
The heat from the channel is carried downstream with the fluid, but a small
fraction might flow to the surface Q̇CH,S . The surface itself transfers heat to

74
3. ESPSS: European Space Propulsion System Simulation

Figure 3.7.: Channel with relevant areas and surfaces for heat flux calculation

the surrounding Q̇rad . There Q̇CC,W and Q̇rad transfer heat by radiation and
convection. For the calculation of Q̇W,CH , Q̇R,CH and Q̇CH,S one solely has to
consider the convective heat transfer. The heat fluxes Q̇W,R and Q̇R,S within the
solid material are based on the conductive heat transfer.
Prior to the listing of the heat fluxes the geometry of the cooling channels are
explained. Hence, the cross-sectional area of the cooling channel is given once
more in Figure 3.7. Additionally, the relevant areas for the heat flux calculation
are entered.
Aint , Arib , Aext and Ach describe the cross-sectional areas of the wall, the cooling
rib, the surface and the cooling channel. They can be calculated as follows:
 
D ti
Aint = π + wch
n 4
2tc + ti + te
Arib = wr
 2 
D te
Aext = π + wch
n 4
Ach = wc · tc

75
3. ESPSS: European Space Propulsion System Simulation

Sw , Sw,r , R, Sch,int , Sch,rib , Sch,ext , Sr,s , and Sext describe the intermediate
surfaces between the internal space of the combustion chamber or nozzle, the
wall, the cooling rib, the cooling fluid, the surface and the surrounding. They can
be calculated as follows:

Sw = Sext = l · (wch + 2wr )


q
Sw,r = l · wr2 + t2i
q
Sr,s = l · wr2 + t2e

Sch,int = Sch,ext = l · wch


Sch,rib = l · tch

By means of these values, the equations for the heat fluxes and the energy
conservation can be set up. In the following the heat fluxes and subsequent the
equations for the energy conservation are listed for a segment i:



 q̇wall = hc Awet (Taw − tp.T ) + σAwet (Tc4 − tp.T 4 )
lπD λint


q̇cha = (tp.T − Tw,int )





 2n ti /2

 lwc λint
q̇coo,int = (Channel.tpin .T − Tw,int )


2 ti /2



lwc λext



q̇coo,ext = (Channel.tpout .T − Tw,ext ) (3.55)
2 te /2
λint

 
q̇coo,rib = lwc Channel.tplat .T − Tw,rib





 wr /2
λint

 
q̇ = S T − T

rib,int w,r w,rib w,int

(ti + tc + wc )/2





 λext 
q̇rib,ext = Sr,s Tw,rib − Tw,ext


(te + tc + wc )/2

where λint and λext are the heat conductivities of the internal walls/ribs and
external walls, respectively. Please note that the heat flux q̇wall equals the heat
flux from the combustion chamber (Eq. 3.39).
Additionally to the radial heat fluxes (Equations 3.55), the longitudinal heat
fluxes need to be regarded. Since the cooling channel has four solid cross-sectional
areas (AS, 2 x AR, AW), four longitudinal heat fluxes are calculated at every
junction. Figure 3.8 illustrates these heat fluxes for a segment i.

76
3. ESPSS: European Space Propulsion System Simulation

Figure 3.8.: Longitudinal heat fluxes for a segment i

The longitudinal heat fluxes are based solely on the heat transfer mechanism
conduction. In the following the associated equations are given:

2λint


 q̇z,int (i) = Aint (Tw,int (i − 1) − Tw,int (i))


 li−1 + li
 2λext
q̇z,ext (i) = Aext (Tw,ext (i − 1) − Tw,ext (i)) (3.56)
 li−1 + li
2λint

 

q̇z,rib (i) = Arib
 Tw,rib (i − 1) − Tw,rib (i)
li−1 + li

In order to calculate the energy conservation of the cooling jacket, one has to
regard the different volumes separately. Subsequent the equations are given for
the wall, the cooling rib and the surface:

dTw,int (i)


 (ρint · Vint )i Cp(i) = q̇cha (i) − q̇coo,int (i) − q̇rib,int (i) + q̇z,int (i) − q̇z,int (i + 1)

 dt
dTw,ext (i)

(ρext · Vext )i Cp(i) = −q̇ext (i) − q̇coo,ext (i) − q̇rib,ext (i) + q̇z,ext (i) − q̇z,ext (i + 1)
 dt
 dTw,rib (i) (ρ · V ) Cp(i) = q̇



int rib i rib,int (i) + q̇rib,ext (i) − q̇coo,rib (i) + q̇z,rib (i) − q̇z,rib (i + 1)
dt
(3.57)

77
4. Steady State Library

The already available propulsion library ESPSS can be used to study both station-
ary states and transients of a propulsion system. Unfortunately its use for steady
state applications is not trivial because of the complexity of the transient models
there implemented. Therefore, simplified pre-design and parametric studies are
difficult and time-consuming. The library shown hereafter is designed specifically
for steady state purposes, providing a helpful and fast tool for the pre-design phase
(feasibility analysis) allowing for parametric studies.
To this aim, the available fluid properties and combustion modelling functions
of ESPSS have been implemented in an adequate form into new libraries. Addi-
tionally, fluid dynamic, combustion and heat transfer models have been developed
to simulate the physical steady state behaviour of the main components of a
propulsion system, as pipes, valves, turbines, pumps, orifices, combustion cham-
ber and nozzle. These components are suited for both launcher and spacecraft
applications.

4.1. Components Overview

The set of models developed for the State state library are able to represent most of
the components of a liquid propulsion system for spacecraft or rockets. Figure 4.1
shows an overview of the components model developed for the steady state
library.

4.2. Ports

Ports are used to connect components to each other, in order to guarantee the
propagation of the variables. Two new ports have been created: the Steady
State fluid port [34], that represents the basis and the rationale on which the

78
4. Steady State Library

Figure 4.1.: Components in the Steady State library

79
4. Steady State Library

whole library is coded, and the nozzle port, to better manage nozzle connections,
similarly to the transient nozzle port.
Each port can connect two or more components at once. SUM variables, such
as mass flow rate, will be summed at the ports to ensure mass flow conservation;
EQUAL variables, such as stagnation pressure, will be propagated to all compo-
nents connected to the same port. A standard flow component should have two
ports, one IN and one OUT, thus defining the mass flow direction. IN ports ensure
calculation of the enthalpy flow ṁh, while enthalpy h is computed in the OUT
ports.
Similarly to the original ESPSS fluid port, the Steady State fluid port can
propagate the molar fraction of chemical species to allow the correct evaluation of
fluid flow of combustion products in systems where this is required (e.g. staged
combustion cycles).

4.3. The “type” switch

Most components have a “type” switch, that enable switching the model between
Design and Off-Design mode.

• Design mode. Geometrical construction data for junctions and valves are
an output of the components. They are calculated from a given ∆P .
The combustion chamber requests chamber pressure, mixture ratio MR and
throat diameter as main design inputs. The propellant mass flows and heat
fluxes are its main outputs.
Turbomachinery components in design mode evaluate performance parame-
ters as torque, power and rotational speed using the mass flow and pressures
coming from the ports.

• Off-Design mode. This mode can be used for the analysis of a given cycle
with fixed geometry and main characteristics. Here, junctions and valves
have a given geometry. Mass flow or ∆P are calculated, depending on the
relative placement of the components.
Combustion chambers have a given nozzle throat diameter (as in design

80
4. Steady State Library

mode), but mass flows are given from the inlet ports. Chamber pressure and
MR are calculated accordingly.

From this general description some additional options are given in Design
mode, in order to fine-tune the models depending on the cycle studied. Therefore,
mass flow ratios through splits can be user-fixed (inputs), or calculated as outputs.
Likewise, turbine pressure ratios can be fixed or calculated from the given cycle.

4.4. 1-D pipes

Tubes components are able to evaluate a one-dimensional flow in steady state


conditions. The tube takes into account the enthalpy variation due to external
heat fluxes and the pressure drop due to the friction along the pipe. As in the
transient version of the component, the steady state tube is divided in volumes
and junctions. Pressure drops and enthalpy variations are calculated at the end of
each volume, on the junction.
The governing equations based on the one-dimensional steady state model are
the following:

Mass conservation
d
(ρvA) = 0 (4.1)
dx

Momentum conservation

dP dv 1 2 fr
+ ρv + ρv =0 (4.2)
dx dx 2 Di

where the friction factor fr is a function of Re and relative roughness, as defined


in Ref. [43].

Energy conservation
dh0
ρvA = hc (Tw − T ) Pw (4.3)
dx

81
4. Steady State Library

Variable Description Unit


Inputs
ṁ Mass flow rate [kg/s]
h1 Inlet enthalpy [kJ/kg]
P2 Outlet Pressure [Pa]
Parameters
Po Initial pressure [Pa]
To Initial temperature [K]
rhoo Initial density [kg/m3 ]
mo Initial mass flow [kg/s]
rug Roughness [m]
kf Multiplier of the friction factor [-]
alphab end Bend angle [deg]
Rb end Ratio of curvature bend [m]
flda dd Additional losses in f L/D [-]
ht_option Heat transfer option [-]
L Tube length [m]
D Nominal tube inner diameter [m]

Table 4.1.: 1-D pipe element

The heat flux is evaluated using the thermal port and connecting the com-
ponent with all components inside the EcosimPro thermal library. It has been
decided to use the original transient EcosimPro thermal library, but still allowing
to be interfaced with the steady state library. This choice enables a relaxation
in the overall steady state model stiffness thanks to the first order differential
equations present in the thermal components. Of course, due to the specifications
of the steady state library, the time variable will have no physical meaning
anymore, and it must rather be regarded as an integration constant.
The pipe component is inherited from the tube. Additionally it features a
1-D wall model for the evaluation of the heat fluxes and heat capacities. As
in the corresponding transient component, it includes a material pipe thermally
connected to the tube, and permits simple convection with the ambient by using a
constant convective coefficient hc .

82
4. Steady State Library

4.5. 0-D components: junctions & valves

0-D components represent concentrated pressure loss in a propulsion system, such


as orifices and valves. These components are based on a similar model; the valve
component differs from the orifice only because it enables the variation of the
throat area while the orifice presents a constant one. Both the components feature
a Design and an Off-Design mode.

Variable Description Unit


Inputs
ṁ Mass flow rate [kg/s]
h1 Inlet enthalpy [kJ/kg]
P2 Outlet Pressure [Pa]
Outputs
h2 Outlet enthalpy [kJ/kg]
P1 Inlet Pressure [Pa]
A Valve Area [m2 ]
Parameters
∆P Pressure loss (Design mode) [Pa]
ζ Pressure drop coefficient [-]
mo Initial mass flow [kg/s]

Table 4.2.: 0-D Junction element

In the Design mode the pressure drop is not related to the mass flow but is an
input coming from the ports or from the user and the geometric parameters of the
junction/valve are calculated:

2∆P ρ 2∆P ρ
Kjun = Kvalve = 2
ṁ2 ṁ pos2
r
ζ
A=
K

In the second case, the Off-Design mode, the model calculates the concentrated
pressure drop as:
ṁ2
∆P = ζ (4.4)
2ρA2

83
4. Steady State Library

and the mass flow is calculated implicitly from Equation 4.4.

4.6. Combustion Chambers

The thrust chamber is composed by two different components following the same
idea developed in the transient library. A combustor component and a nozzle
component are linked together to create the thrust chamber.

Combustor

This component represents a non adiabatic 1-D combustion process inside a


convergent chamber (up to the throat section). It is a steady state, one dimensional,
isoenthalpic combustion chamber. The equilibrium combustion products are
calculated using the minimum Gibbs energy method [55] as a function of the
propellant’s mixture molar fractions and enthalpies and of the chamber pressure.
The chamber geometry allows for precise chamber contour definitions and non
homogeneous node discretisation.
Thermodynamic properties along the chamber sections are evaluated using
isoenthalpic correlations in frozen conditions. Heat fluxes are calculated with
the Bartz correlation in closed form [9]. The chamber works only in “ignited”
mode, as it is not required for a steady state model to show transitions between
non burning and burning state. The compositions of the combustion products
are evaluated from the injected fluids. It is possible to use either pure fluids or
combustion products from a previous combustor (preburner). Along with the
Design/Off-Design type switch, another switch is responsible for choosing the
combustor type, which can be either a Main Combustion Chamber (MCC) or a
Preburner (PB_GG). The main difference between the two combustor types is the
following:

- MCC. For a Main Combustion Chamber in Design mode, chamber pressure


is user given (Pc = Pdesign )

- PB_GG. For a Preburner in Design mode, chamber pressure is taken from


the outlet port (Pc = f _out.P )

84
4. Steady State Library

Variable Description Unit


Inputs
hox Oxidiser Inlet enthalpy [kJ/kg]
hf u Fuel Inlet enthalpy [kJ/kg]
Outputs
ṁ Mass flow rate [kg/s]
ṁox Oxidiser mass flow rate [kg/s]
ṁf u Fuel mass flow rate [kg/s]
Pc Chamber pressure [Pa]
Tc Chamber temperature [K]
Tw Chamber wall temperature [K]
Parameters
Pco Chamber pressure [initial if Off-D; assigned if D] [Pa]
Tco Initial combustion temperature [K]
Tco x Initial Oxidiser combustion temperature [K]
Tcf u Initial Fuel combustion temperature [K]
MRo Mixture Ratio [initial if Off-D; assigned if D] [-]
ηc Combustion efficiency [kg/s]
Lc Chamber length of subsonic part [m]
Dt Nozzle throat diameter [m]

Table 4.3.: Combustor element

85
4. Steady State Library

In Design mode, the combustor component calculates inlet mass flows from
given chamber pressure, MR and nozzle throat diameter. The mass flow conserva-
tion is written as:

ρth vth Ath


ṁ = (4.5)
ηc
where the subscript th refers to throat conditions, and ηc is the combustion
efficiency.
In Off-Design mode, the component evaluates the equilibrium composition in
the first section to obtain thermodynamic and transport properties, and in the
throat to evaluate the chamber pressure and the mass flow rate by an iterative
loop. This equation is actually used in the overall loop to determine the chamber
pressure Pc implicitly. The ideal gas equation is written twice, for stagnation
chamber and for throat conditions. Isentropic throat conditions are calculated
iteratively assuming shifting equilibrium and variable isentropic coefficient γ .
For each node i, the relevant characteristics (Mach number Mi , Pi , Ti , ρi ,
sound speed vsound,i , vi ) are calculated assuming isentropic flow conditions. This
simplification is acceptable since these variables are only needed for assessing the
heat transfer coefficient within the Bartz equation.

Nozzle

The component represents a 1-D supersonic nozzle in steady state conditions.


The choked throat conditions (Pth , Tth and vth ) are evaluated in the combustor
component and communicated through the nozzle port. Stagnation conditions
are calculated from the throat conditions. Static conditions are evaluated in each
section using isentropic correlations and assuming frozen chemistry. The heat
flux in each section is evaluated using the semi-empirical correlation of Bartz in a
closed form.

86
4. Steady State Library

Variable Description Unit


Inputs
h Enthalpy at throat [J/kg]
s Entropy at throat [J/kg K]
P Pressure at throat [Pa]
T Temperature at throat [K]
ρ Density at throat [kg/m3 ]
Cp Specific heat at throat [J/ kg k]
ṁ Mass flow rate [kg/s]
Outputs
Pi Nozzle pressure profile [Pa]
Ti Nozzle temperature profile [K]
Tw Nozzle wall temperature [K]
Isp Specific impulse [s]
Thrust Thrust [N]
Parameters
Pco Initial nozzle pressure [Pa]
Tco Initial nozzle temperature [K]
ηCf Nozzle efficiency [-]
Pext External pressure [Pa]
Ld Nozzle length of supersonic part [m]
Dt Nozzle throat diameter [m]

Table 4.4.: Nozzle element

87
4. Steady State Library

4.7. Cooling Channels

Cooling channels component features two different models depending on the


complexity of the propulsion system:

• cooling jacket

• regenerative circuit

The cooling jacket component is inherited from the tube component. It permits
the modelling of a combustion chamber cooling jacket. Mass flow, pressure drop
and temperature rise of the coolant are evaluated using the same governing
equations as for the pipes and the tubes (see Equations 4.1, 4.2, 4.3). Since the
direction of the flow in the Steady State library must be given at the schematic
design stage, two components are foreseen: a co-flow and a counterflow cooling
jacket. They are constructed by aggregation of one tube representing the channels
and a simplified 3D geometry (built by means of several bars around the channels)
around them (see Figure 4.2). The regenerative circuit component features a
pre-defined pressure drop and a pre-assigned hot gas side wall temperature Tw,hot
profile along the combustion chamber. Mass flow value is coming from the ports;
the component sends the Tw,hot variable to the combustion chamber through the
thermal port. In this way it is possible to evaluate the chamber heat flux q̇w (see
Eq. 4.6). Using this variable, the enthalpy rise along the channel is evaluated and
so the outlet temperature. Choosing a material for the chamber wall, the chamber
wall thickness is an output of the design (see Eq. 4.7). The channel height is a user
given input for the model, while the channel width is evaluated from the fixed
number of channels.

q̇w = hc Awet (Taw − Tw,hg ) (4.6)


λ
q̇w = Awet (Tw,hg − Tw,cf ) (4.7)
t

88
4. Steady State Library

Variable Description Unit


Inputs
ṁ Mass flow rate [kg/s]
h1 Inlet enthalpy [kJ/kg]
P2 Outlet pressure [Pa]
Outputs
P1 Inlet pressure [Pa]
h2 Outlet enthalpy [kJ/kg]
T2 Outlet temperature [K]
Tcool Coolant temperature [K]
Tw Channel wall temperature [K]
q̇w Chamber heat flux [W]
Parameters
nch Number of channels [-]
Po Initial pressure [Pa]
To Initial temperature [K]
rhoo Initial density [kg/m3 ]
mo Initial mass flow [kg/s]
rug Roughness [m]
kf Multiplier of the friction factor [-]
alphab end Bend angle [deg]
Rb end Ratio of curvature bend [m]
flda dd Additional losses in f L/D [-]
ht_option Heat transfer option [-]
wch Channel widths [m]
tch Channel heights [m]
thi Jacket inner wall thickness [m]
the Jacket outer wall thickness [m]
mati Jacket internal material [-]
mate Jacket external material [-]
L Channel length [m]
D Nominal channel inner diameter [m]

Table 4.5.: Cooling jacket element

89
4. Steady State Library

Variable Description Unit


Inputs
ṁ Mass flow rate [kg/s]
h1 Inlet enthalpy [kJ/kg]
P2 Outlet pressure [Pa]
Outputs
P1 Inlet pressure [Pa]
h2 Outlet enthalpy [kJ/kg]
T2 Outlet temperature [K]
Tcool Coolant temperature [K]
q̇w Chamber heat flux [W]
tw Internal wall thickness [m]
Parameters
nch Number of channels [-]
Po Initial pressure [Pa]
To Initial temperature [K]
rhoo Initial density [kg/m3 ]
mo Initial mass flow [kg/s]
dPdesign Design pressure drop [Pa]
Tw Channel wall temperature [K]
wch Channel widths [m]
tch Channel heights [m]
the External wall thickness [m]
mati Jacket internal material [-]
mate Jacket external material [-]
L Channel length [m]
D Nominal channel inner diameter [m]

Table 4.6.: Regenerative circuit element

90
4. Steady State Library

Figure 4.2.: Cooling jacket channels wall mesh [43]

4.8. Turbomachinery

4.8.1. Pump

The pump component features a simple model using isentropic relations and
constant, user-given efficiency ηp to calculate pump conditions. The isentropic
enthalpy rise is calculated assuming an isentropic transformation between inlet
and outlet pressure.
The Design type parameter decides whether the pump pressure rise is fixed
from the ports (in Design mode) or calculated (in Off-Design mode). In both
modes, given a assigned specific speed Ns, the shaft rotation speed ω is calculated:
p
ω Q/ns
Ns = (4.8)
(TDH/nst )0.75

where ns is the number of suctions, nst is the number of stages of the pump, Q
is the volumetric flow and TDH represents the actual total dynamic head of the
pump. In Design mode, the inlet entropy s is calculated from inlet pressure Pin
and enthalpy hin . Then, the isentropic enthalpy rise dhis is calculated with an
isentropic transformation, with known inlet and outlet pressures Pin and Pout ,

91
4. Steady State Library

Variable Description Unit


Inputs
h1 Inlet enthalpy [J/kg]
P1 Inlet pressure [Pa]
T1 Inlet temperature [K]
ṁ Mass flow rate [kg/s]
Outputs
P2 Outlet pressure [Pa]
ω Rotational speed [rad/s]
τ Torque [N m]
W Power [W]
Parameters
Po Initial pressure [Pa]
To Initial temperature [K]
mo Initial mass flow [kg/s]
rpmo Initial rotational speed [rad/s]
ηp Pump efficiency [-]

Table 4.7.: Pump element

and inlet entropy sin . Finally, the real enthalpy rise dh is calculated as:

dhis
dh = (4.9)
ηp

and subsequently, the shaft rotational speed ω and torque τ are then linked with
the actual enthalpy rise by the power balance equation:

ω τ = ṁ ∆h (4.10)

In Off-Design mode, the inlet entropy sin is calculated (as for the Design mode)
from inlet pressure Pin and enthalpy hin . The real enthalpy rise dh is given by the
power balance equation, and the isentropic enthalpy rise dhis is given by:

dhis = η · dh (4.11)

Therefore, the outlet pressure Pout is calculated with an isentropic transformation,


with known isentropic enthalpy rise dhis and entropy s. As for the Design case,

92
4. Steady State Library

it is possible to compute torque and power from it (thank to an assigned specific


speed Ns).
In the near future the off-design mode (with assigned specific speed) will
be upgraded with the capability of using performance maps for efficiency and
pressure head.

4.8.2. Turbine

The turbine component is a model using isentropic relations and constant, user-
given efficiency ηt to calculate turbine conditions. The isentropic enthalpy fall is
calculated assuming an isentropic transformation between inlet and outlet pressure.
The construction parameter turbine_type decides, in Design mode, whether the
mass flow ṁ is assigned from the ports (known_mflow) thus calculating the
upstream pressure, or the pressures are assigned from the ports (known_pressures,
known_PI_tt), thus calculating the mass flows. This switch must be carefully set
depending on the cycle studied:

• turbine_type = known_mflow. For closed cycles, in most cases, the mass


flow is determined by the preburner (staged combustion) or by bypass valves
(expander), therefore it is given from the ports, and the pressure ratio should
be calculated in the turbine component.

• turbine_type = known_pressures. For open cycles (gas generator) or for


one of the turbines in closed cycles with parallel turbines, the pressures
should be fixed (known_pressures), and the model will find the mass flow
that equilibrates the pump power.

• turbine_type = known_PI_tt. For open cycles, especially in the pre-design


phase to have a first attempt result and to facilitate parametric studies
changing the pressure ratio Πtt ; the model will find the mass flow that
equilibrates the pump power.

In all working modes (Design or Off-Design, known_mflow or known_pressures,


known_Pi_tt), torque, shaft rotation speed ω and power are given from the
mechanical port.

93
4. Steady State Library

Variable Description Unit


Inputs
h1 Inlet enthalpy [J/kg]
P2 Outlet pressure [Pa]
T1 Inlet temperature [K]
W Power [W]
ṁ Mass flow rate (if know_mflow) [kg/s]
P1 Inlet pressure (if known_pressures) [Pa]
PItt Pressure ratio (if known_PI_tt) [-]
Outputs
ṁ Mass flow rate (if know_pressures or PI_tt) [kg/s]
P1 Inlet pressure (if know_mflow) [Pa]
ω Rotational speed [rad/s]
τ Torque [N m]
Parameters
Po Initial pressure [Pa]
To Initial temperature [K]
mo Initial mass flow [kg/s]
PI_tto Initial pressure ratio [-]
rpmo Initial rotational speed [rad/s]
ηt Turbine efficiency [-]

Table 4.8.: Turbine element

94
4. Steady State Library

• Design mode

- known_mflow. The mass flow is used to calculate the real enthalpy


fall (Eq. 4.12).
dh = −P ower/ṁ (4.12)

From this value, the isentropic enthalpy fall is calculated with the fixed
efficiency η (Eq. 4.13).
dhis = dh/η (4.13)

The inlet entropy sin is calculated from the outlet pressure Pout and
the ideal outlet enthalpy hout = hin − dhis . Then, the inlet pressure is
calculated from entropy sin and inlet enthalpy hin .
- known_pressures. The inlet entropy sin is calculated from the inlet
pressure Pin and the inlet enthalpy hin . Then, the isentropic enthalpy
fall dhis is calculated knowing the inlet enthalpy hin , the outlet pressure
Pout and the inlet entropy sin (Eq. 4.14).

dhis = hin − f (sin , Pout ) (4.14)

Thereafter, the real enthalpy fall is given by using the turbine efficiency
η (Eq. 4.15).
dh = η · dhis (4.15)

The mass flow is finally estimated from Eq. 4.12.


- known_PI_tt. When this turbine type is chosen, the calculation proce-
dure differs from the one used in known_pressures only by the inlet
pressure Pin given as an output from Pout times Πtt .

• Off-Design mode. The procedure is roughly the same as the Design mode
case with turbine_type = known_pressures.

4.9. Validation

Several test cases have been performed in order to evaluate the reliability of the
Steady State library. Following a step by step approach, first each component

95
4. Steady State Library

singularly and then more complex systems were validated.

4.9.1. Component validations

The schematics shown hereafter are only the graphical interfaces of mathematical
models where all variables of each component are considered. The tool is able of
calculating the steady state of the mathematical model by solving the non-linear
algebraic equation system that results from the built schematic by means of the
“Newton-Raphson” or the “Minpack” method [40].
In order to have a flexible and robust tool, each component model is carefully
coded to provide the most suitable variables that can be used to break the non-
linear algebraic loops. The choice of the correct variables that enable the equations
system to converge represents one of the most important achievements of this
work.

Pipeline test case

The purpose of this test case is to validate the Steady State pipe component
and demonstrate its proper function compared to a transient component. The
schematic shown in Figure 4.3 has been also built to check the correct behaviour
of Steady State components in long pipelines. A long pipeline is modelled twice,
with standard ESPSS transient components and with Steady State components.
Pressure drop distribution along the pipeline and mass flow rates are compared
between the two models.
Please note the absence of the volume between two junctions. Purely capacitive
components are not needed in the Steady State library, and it is possible to chain
multiple resistive components in series.
The schematic describes a series of pipes linked together by junctions. A
pressure difference has been imposed between inlet and outlet. The input data
shown in Table 4.9 represent the inputs implemented in each component (steady
or transient). The initial conditions have been taken equal for each pipe, with
atmospheric pressure and low initial mass flow. These conditions are quite distant
from the solution, and the convergence of the steady state code in this case is an
indicator of its robustness.

96
4. Steady State Library

Figure 4.3.: Schematic of the Pipeline test case. Purple: steady state components.
Cyan: transient components

Name Description Value Units


Pin Total Pressure at inlet 50 [bar]
Tin Total Temperature at inlet 300 [K]
Pout Total Pressure at outlet 30 [bar]
Po Initial Total pressure in the pipe 1 [bar]
To Initial Total temperature in the pipe 300 [K]
mo Initial mass flow in the pipe (guess value) 0.2 [kg/s]
rug Roughness 5e-05 [m]
L Pipe length 1 [m]
D Pipe internal diameter 0.01 [m]
nodes Pipe nodes discretisation 5 [-]
Ao Junction area 7e-05 [m2 ]
ζ Loss coefficient 1 [-]
fluid Working fluid Real H2O [-]

Table 4.9.: Pipeline input data

97
4. Steady State Library

The simulation results are summarized in Table 4.10, showing a very accurate
mass flow calculation and pressure drop distribution.

Name Value Transient Value Steady State Error


m [kg/s] 1.109 1.109 0.006%
∆P1 [bar] 3.110 3.110 0.002%
∆P2 [bar] 3.111 3.111 0.002%
∆P3 [bar] 3.112 3.112 0.002%
∆P4 [bar] 3.112 3.112 0.002%

Table 4.10.: Pipeline output data

Combustion Chamber

The purpose of this test case is to calculate the main characteristics of combustion
chamber and nozzle components. Its schematic is shown in Figure 4.4 for both
steady state and transient models. The only difference between the two models
(besides the different modelling approach) is the absence of the injector capacity
inside the Steady State combustion chamber injection plate.
Relevant input data are listed in Table 4.11. The initial conditions are the same
for the steady state and the transient components.
The test compares the propellant mass flows and the chamber pressure and
temperatures between the two models. Other important characteristics as heat
fluxes, wall temperatures and adiabatic wall temperatures have been evaluated as
well, but are not reported here for simplicity.
The output data from the two models are compared in Table 4.12. It is evident
that the steady state results are very similar to the respective transient results.

98
4. Steady State Library

Figure 4.4.: Schematic of Combustion Chamber test case. Purple: steady state
components. Cyan: transient components

Name Description Value Units


Pin,ox Ox. Total Pressure at inlet 70.8 [bar]
Tin,ox Ox. Total Temperature at inlet 94.7 [K]
Pin,f u Fu. Total Pressure at inlet 71 [bar]
Tin,f u Fu. Total Temperature at inlet 208.9 [K]
Nsub Number of subsonic nodes 5 [-]
Nsup Number of supersonic nodes 5 [-]
Lcc Chamber length of subsonic part 0.5 [m]
Dth Nozzle throat diameter 0.10 [m]
Pcc Initial Chamber pressure 1 [bar]
Tcc Initial Chamber temperature 300 [K]

Table 4.11.: CC input data

99
4. Steady State Library

Name Transient Value Steady State Value Error


mox [kg/s] 18.72 18.53 1.0 %
mf u [kg/s] 3.14 3.11 1.0 %
mtot [kg/s] 21.86 21.64 1.0 %
MR [-] 5.959 5.956 0.04 %
Pcc [bar] 64.97 64.27 1.1 %
Tcc [K] 3518 3514 0.11 %
Mach [-] 2.887 2.762 4.5 %

Table 4.12.: CC output data

4.9.2. Subsystem validations

HM7B Turbopump subsystem

This test case was used during the ESPSS Industrial Evaluation from Astrium
Bremen to validate the ESPSS library for liquid rocket engine cycles [41].
The schematic shown in Figure 4.5 represents the turbomachinery power pack
of the upper stage engine of the Ariane 5 launcher, the HM7B engine, including
the gas generator and both turbopumps.
Figure 4.6 shows the equivalent schematic implemented with Steady State
components. They are very similar to each other. Only volume components and
non-condensable fluid lines are absent. The first ones are not needed for the same
reasons stated in Section 4.9.1; the latter have been eliminated since there is no
need to model the Helium purging phases in a steady state simulation. The steady
state model has been used in Off-Design mode.
The chosen input data are collected in Table 4.13; Table 4.14 summarizes the
main system variables results performed by the transient and the steady state
models. The steady state model matches very well the transient one for fluid flow,
turbomachinery and gas generator main parameters.

100
4. Steady State Library

Name Description Value Units


Pin,ox Total pressure in LOX tank 2.0 [bar]
Pin,f u Total pressure in LH2 tank 3.0 [bar]
Pout,ox Total pressure at Pump outlet/Gas Generator inlet 50.0 [bar]
Pout,f u Total pressure at Pump outlet/Gas Generator inlet 55.0 [bar]
Pcc Initial chamber pressure 20.0 [bar]
Tcc Initial chamber temperature 900 [K]
ωp,ox Initial LOX pump speed 1000 [rpm]
ωp,f u Initial LH2 pump speed 6000 [rpm]

Table 4.13.: HM7B Turbopump input [44] and initial data

Name Nominal Value [44] Error Transient/Steady State


ṁgg,ox [kg/s] 0.3%
ṁgg,f u [kg/s] 0.07%
M R [-] 0.5%
Pgg [bar] 1.3%
Tgg [K] 0.08%
ṁt [kg/s] 0.3%
ωt [rpm] 60500 0.8%
τt [N·m] 59.98 2.0%
ṁp,ox [kg/s] 12.4 0.0%
∆Pox [bar] 48 1.6%
ωp,ox [rpm] 13000 0.8%
τp,ox [N·m] 10.4%
ṁp,f u [kg/s] 2.4 0.0%
∆Pf u [bar] 52 0.8%
ωp,f u [rpm] 60500 0.8%
τp,f u [N·m] 6.0%

Table 4.14.: HM7B Turbopump output data

101
4. Steady State Library

Figure 4.5.: Turbopump test case: HM7B power pack transient schematic

Figure 4.6.: Turbopump test case: HM7B power pack steady state schematic

102
4. Steady State Library

HM7B Chamber subsystem

This test case represents the combustion chamber subsystem of the HM7B engine.
The aim of this test case is to validate the behaviour of the combustion chamber
and cooling jacket components when they are coupled together in a simulation,
by comparing results with the transient model simulation.
Figures 4.7 and 4.8 show the schematics of the combustion chamber subsystem
using the ESPSS transient library and the steady state model, respectively.
As in the previous test case, described in Section 4.9.2, the similarity of the two
schematics shown hereafter is evident. The only difference for the steady state
model is the absence of non-condensable fluid lines and capacitive components
such as volumes. Also here the steady state model has been used in Off-Design
mode.

Figure 4.7.: Chamber test case: HM7B Combustion Chamber transient schematic

In Table 4.15 the main input data for both systems are collected; in Table 4.16
the main system variables results are summarized, performed by the transient and
the steady state models. As reported in the table the steady state model matches
the transient results, showing very good agreement between the values of the
combustion chamber, and of the cooling channel model.

103
4. Steady State Library

Figure 4.8.: Chamber test case: HM7B Combustion Chamber steady state
schematic

Name Description Value Units


Pin,LOX Total pressure at pump outlet/chamber inlet 50.0 [bar]
Pin,H2 Total pressure at pump outlet/chamber inlet 55.0 [bar]
Pcc Nominal chamber pressure 36.6 [bar]
nch Numbers of channels 128 [-]
Pi,cc Initial chamber pressure 30 [bar]
Ti,cc Initial chamber temperature 1000 [K]
Po Initial total pressure in the channels 49 [bar]
To Initial total temperature in the channels 30 [K]
mo Initial mass flow in the channels 2 [kg/s]

Table 4.15.: HM7B CC input [44] and initial data

Name Nominal Value [44] Error Transient/Steady State


mox [kg/s] 12.4 2.4%
mf u [kg/s] 2.46 0.38%
mtot [kg/s] 14.86 2.0%
MR [-] 5.0 2.9%
Pcc [bar] 36.6 0.32%
Tcc [K] 0.88%
mch [kg/s] 2.46 0.37%
∆Pch [bar] 9.9%
Tout,ch [K] 6.7%

Table 4.16.: HM7B CC output data

104
4. Steady State Library

4.9.3. Engine cycle designs

At the beginning of a design analysis, a set of performance parameters must be


chosen as assumption to define the engine class and the initial condition of the
engine:

- Propellants

- Tank pressure and temperatures

- Chamber Pressure P c

- Chamber Mixture Ratio MR

- Combustion efficiency ηc∗

- Throat Diameter Dt

- Pump efficiencies ηp

- Pump specific speeds Ns

- Turbine efficiencies ηt

Subsequently, it is possible to evaluate other important characteristics such as


the contour of the chamber (by use of L* and simple geometrical correlations) and
injector pressure drops.

HM7B rocket engine system

The HM7B is a gas generator cycle with single turbine and geared pumps. The
two subsystems modelled in the previous work have been updated to the current
library implementation and linked together to build the HM7B engine system
model.
Figure 4.9 shows the schematic of the HM7B engine using the Steady State
library. All model components are in Design mode.
In Table 4.17 the main input data for the system are collected; in Table 4.18 the
main system variables results are summarized and compared with nominal data.
Where nominal values are shown, they are taken from the open literature.

105
4. Steady State Library

Figure 4.9.: HM7B engine system schematic

106
4. Steady State Library

Name Description Value Units


Pin,LOX Total pressure in LOX tank 2.0 [bar]
Tin,LOX Total temperature in LOX tank 91.2 [K]
Pin,LH2 Total pressure in LH2 tank 3.0 [bar]
Tin,LH2 Total temperature in LH2 tank 21.0 [K]
Pcc Nominal chamber pressure 36.6 [bar]
MR Nominal chamber mixture ratio 5. [-]
nch Numbers of channels 128 [-]
Ti,cc Initial chamber temperature 1000 [K]
Po Initial total pressure in the channels 50 [bar]
To Initial total temperature in the channels 30 [K]
mo Initial mass flow in the channels 2 [kg/s]
Pc,gg Initial gas generator pressure 20 [bar]
Tch Initial gas generator temperature 900 [K]
rpmox Initial LOX pump speed 1000 [rpm]
Ns,ox LOX pump specific speed 10.95 [-]
rpmf u Initial H2 pump speed 6000 [rpm]
Ns,f u H2 pump specific speed 9.29 [-]
mo,tu Initial Turbine mass flow 0.2 [kg/s]

Table 4.17.: HM7B input [44] and initial data

Pressure drop has been fixed in each valve and junction as well as turbomachin-
ery efficiency and gas generator mixture ratio. For the turbine, the “known_PI_tt”
type is chosen, while the pump specific speed Ns has been calculated from design
data.
The main chamber pressure and mixture ratio are fixed. Propellant mass flows
to the main chamber are calculated, and fed back to the upstream components.
The gas generator mass flow is not fixed. Only its mixture ratio is fixed, in
accordance with the maximum allowable temperature in the turbine. The mass
flow is then calculated by an algebraic equation system resulting automatically
from the connection of gas generator and turbopumps. The needed shaft power
drives the total gas generator mass flow rate, since the turbine pressure ratio is
fixed by design.
Initial values such as mass flow rates and shaft speed are chosen by rough engi-
neering assessments. The robustness of the Steady State library is demonstrated

107
4. Steady State Library

Name Nominal Value [44] Error


ṁox,gg [kg/s] 0.59%
ṁf u,gg [kg/s] 0.23%
ṁt,gg [kg/s] 0.3%
Pgg [bar] 0%
Tgg [K] 1.03%
mcc,ox [kg/s] 0.88%
mcc,f u [kg/s] 0.83%
mcc,t [kg/s] 14.86 0.87%
mch [kg/s] 2.36%
∆Pch [bar] 0.5%
Tout,ch [K] 6.7%
ωt [rpm] 60500 0.16%
τt [N·m] 4.48%
Wt [W] 4.32%
Tin,t [K] 11.3%
ṁp,LOX [kg/s] 0.96%
∆PLOX [bar] 48 0.41%
ωp,ox [rpm] 13000 0.16%
τp,ox [N·m] 1.39%
ṁp,LH2 [kg/s] 7.56%
∆PLH2 [bar] 52 1.29%
ωp,f u [rpm] 60500 0.16%
τp,f u [N·m] 6.0%

Table 4.18.: HM7B engine system output data

108
4. Steady State Library

by the stability of the simulation in a wide range of initial conditions.


From Table 4.18 a very good agreement with nominal data is recognisable. Few
parameters have an higher percentage error: the fuel mass flow rate in the pump
does not take into account the dump and the tap-off mass flow rate vented from
the engine. The cooling jacket exit temperature takes into account the temperature
increase in the injector dome. If compared to the LH2 injector dome temperature,
the error decreases to 2.47%. Turbomachinery parameters show quite good results;
the differences are mainly due to the turbine inlet temperature that is lower then
the nominal one.

109
4. Steady State Library

RL-10A-3-3A rocket engine system

The RL-10 is an expander cycle with single turbine and geared pumps. Being a
closed cycle, the cycle design is more difficult than for an open cycle, because all
parameters are strongly dependent from each other.
In this test case a model of the RL-10A-3-3A rocket engine system has been
created and simulated in Design mode. Model results from the steady state
calculations have been compared with the typical engine performance parameters
at nominal operating conditions [15, 4].
Pressure drops of the main valves and junctions are taken from engine typical
values [15] as well as for the tank pressure and temperature conditions. Pumps and
turbine efficiency are fixed to a constant value and taken from open literature [15,
4].
No calibration has been adopted for the control valves: the Oxidiser Control
Valve (OCV) aperture ratio has not been trimmed because the mixture ratio
is assigned in the combustion chamber. Mass flows are given by combustion
chamber conditions. Turbomachinery power is regulated by the Thrust Control
Valve (TCV) that is open at its nominal open area ratio of 9% [58]. In Table 4.19 the
main input data for the system is collected; in Table 4.20 the main system variables

Name Description Value Units


Pin,LOX Total pressure in LOX tank 2.43 [bar]
Tin,LOX Total temperature in LOX tank 97.056 [K]
Pin,LH2 Total pressure in LH2 tank 1.86 [bar]
Tin,LH2 Total temperature in LH2 tank 21.44 [K]
Pcc Nominal chamber pressure 32.75 [bar]
MR Nominal chamber mixture ratio 5.055 [-]
nch Numbers of cooling channels 180 [-]
Ti,cc Initial chamber temperature 1000 [K]
mo Initial mass flow in the channels 2 [kg/s]

Table 4.19.: RL-10A-3-3A input and initial data

results are summarized, performed by the steady state model and compared with

110
4. Steady State Library

Figure 4.10.: Schematic of the RL-10 engine

111
4. Steady State Library

performance parameters at nominal condition.

Name Nominal Value Error


mcc,ox [kg/s] 13.95 0.38%
mcc,f u [kg/s] 2.76 0.38%
mcc,t [kg/s] 16.71 0.38%
mch [kg/s] 2.76 0.38%
Tout,ch [K] 213.44 0.18%
ωt [rpm] 31 537 0.39%
ṁt [kg/s] 2.69 1.1%
Wt [kW] 588.36 3.85%
ṁp,LOX [kg/s] 13.95 0.38%
ωp,ox [rpm] 12 948 0.4%
Wp,ox [kW] 82.026 4.07%
ṁp,LH2 [kg/s] 2.7946 0.85%
ωp,LH2 [rpm] 31 537 0.39%
Wp,LH2 [kW] 501.86 3.68%
T hrust [kN] 73.4 2.89%
Isp [s] 444 2.5%

Table 4.20.: RL-10A-3-3A engine system output data

The chamber pressure is assigned together with the mixture ratio. The pressure
cascade for both propellant lines is calculated by the model according to design
parameters such as valve pressure drops. The pressure rise for the LOX pump
results directly from the calculated LOX pressure cascade, whereas an algebraic
loop is solved for calculating the needed H2 pump pressure rise and the turbine
pressure ratio. In parallel, several other algebraic loops are solved, determining
key variables such as cooling channel outlet temperature and turbine mass flow.
The nominal fuel bypass flow ratio [58] has been fixed using the split component.
The turbine evaluates the required pressure ratio and the mass flow is then
evaluated via the Thrust Control Valve (TCV).
The cooling channel component evaluates iteratively with the combustion
chamber component the heat fluxes and the wall temperatures. Finally, pumps

112
4. Steady State Library

power is evaluated from the required pressure rise and mass flow, and rotation
speed from the given specific speed Ns.
The RL10 design model here described has proven to be very robust with respect
to initial conditions. The comparison of initial conditions in Table 4.19 and results
in Table 4.20 demonstrates this assertion. For example, the initial mass flow in the
cooling jacket mo is 2 kg/s, whereas the simulation result yields 2.76 kg/s.
From Table 4.20 a very good agreement with nominal values is recognisable.

113
5. Transient Modelling

In this chapter we would like to introduce the new models developed for a better
assessment of the phenomena occurring in the subsystems of liquid rocket engines
during start-up. Three new, more complex and accurate models will be presented
in this chapter: the first one for the injector plate, a second one for the evaluation
of the heat transfer coefficient on the hot gas side of the thrust chamber, and the
third one for the evaluation of the thermal stratification inside high aspect ratio
cooling channels (HARCC).

5.1. Injector Plate model

The injector head’s main task is to merge, mix and atomize the oxidizer coming
from the main valve with the fuel coming from the cooling channels. Figure 5.1
shows a schematic illustration of an arbitrary injector head, in order to clarify a
common structure. At first both propellants enter separate volumes in which they
are uniformly distributed among the injector elements. Afterwards the injector
elements dose the amount of propellant mass flow, by a defined pressure drop and
atomize the propellants.
For the computer model and for the mathematical model formulations, respec-
tively, the configuration of both propellant lines in the injector head are simplified
in that way, that one volume and one orifice are assumed in each line. The vol-
ume represents a collector where the propellant is distributed among the injector
elements. Following, the orifices represent the in- and outlet of all the injector
elements. In order to calculate the right flow velocity and Reynolds number in
the injector element, one has to take the associated mass flow rate into account.
Figure 5.2 (a) shows the component wise connection of the above mentioned
volumes and orifices. Additionally, the convective and radiative heat transfer from
the combustion chamber into the injector head has to be regarded. In this process

114
5. Transient Modelling

Figure 5.1.: Schematic illustration of an arbitrary injector head

a heat flux is transferred from the combustion chamber into the fuel cavity first
and subsequently into the oxidizer cavity.
The original injector plate model present in the ESPSS library features a very
simplified thermal model. Indeed the injector plate topology takes into account
the effects of the radiative heat transfer, but the conductive and convective heat
fluxes are evaluated using only a virtual conductance. As in Figure 5.2 (a) the
original injector plate is built by a radiative and a conductive component linked
upstream in parallel directly to the combustor hot gases. These two components
are linked to a capacitive component, used to simulate the total thermal inertia
of the injector cavity walls. This heat capacity is then connected to the two fluid
cavities.
The original model is here described:

q̇cond = λc,hg (Tcore − Tcap )


4
q̇rad = σ (Tcore 4 )
− Tcap


q̇ = q̇ + q̇ + q̇ + q̇cav,f u
cap cond rad cav,ox

115
5. Transient Modelling

(a) topology schematic of the original injector plate

(b) topology schematic of the new injector plate

Figure 5.2.: Schematics of the injector plates

116
5. Transient Modelling

where heat flux of the capacitive component is

dTcap
q̇cap = CpM
dt

and for the cavities components

q̇cav,ox = hc (Tw,cav − Tcav )


q̇cav,f u = hc (Tw,cav − Tcav )

In this way it is not possible to evaluate the presence of convective heat transfer
on injector face plate and to take into account the correct effect of the conductive
and capacitive behaviour of the injector plate material. Moreover, the core tem-
perature in the first volume of the chamber is considered as the wall temperature
of the injector plate, and this is unrealistic.
For this reason, an upgraded version of the injector plate topology inside the
ESPSS library has been implemented [35]. The aim of this new model is to take
into account the convective and radiative heat transfer between the fluid in the
first volume of the chamber and the face plate, and evaluate the conductive and
capacitive effect of the injector walls in more accurate way, representative of
a generic injector head. The new structure of the injector plate (see Figure 5.2
(b)) wants to maintain the level of simplicity of the original model in order to
keep the computational cost low, and to be applicable to several different injector
geometries (impinging, coaxial, etc. . . ), but in the same time wants also to improve
the heat transfer characteristics from the chamber to the injector cavities. In the
first volume of the chamber the convective and radiative heat fluxes to the injector
face are evaluated:

q̇conv = hc,hg (Taw − Tw,hg )


4 4
q̇rad = σ (Tcore − Tw,hg )

using the mass and the material properties of the injector plate (heat capacity, ther-
mal conductivity) the model evaluates the conductive heat transfer and capacity
effect of the walls:

117
5. Transient Modelling

 
λ
q̇cond |ox,f u = (Tw,hg − Tw,cav )
tox,f u
and for the capacitive components

q̇cap,hg = q̇cond,ox + q̇cond,f u + q̇conv + q̇rad


q̇cap,ox = q̇cond,ox + q̇cav,ox


q̇
cap,f u = q̇cond,f u + q̇cav,f u

dT
q̇cap,k = CpMk with k = hg, ox, f u;
dt
The thermal conductivity and heat capacity values are function of the chosen
material of the injector plate and of its temperature. For simplicity reasons the
injector plate is assumed to be made of only one material. The thickness t used for
the evaluation of the conductive heat flux has to be considered as a “characteristic”
injector head thickness or width, and presents two different values, one for the ox
side and the other for the fuel side. The capacitive components for each propellant
side are divided in two parts in order to obtain three different temperatures: Tw,hg ,
Tox,cav , Tf u,cav , respectively the temperature of the injector plate on the hot side,
the oxidizer and the fuel cavity wall temperature on the cold fluid side.

5.1.1. Qualitative behaviour

In order to validate the behaviour of the new injector plate model, a numerical
approach has been used because no experimental results were found in open
literature. A pressure fed propulsion system has been modelled and tested with
both injector plate models. The test case represents a typical spacecraft propulsion
system supplied by nitrogen tetroxide (NTO) as oxidiser, and monomethylhy-
drazine (MMH) as fuel. The system is designed in order to reach a chamber
pressure of ≈ 10 bar with a mixture ratio of 1.65 at steady state conditions.
Figure 5.3 compares the two models by assessing the thermal behaviour inside
the injector cavities and the injector plate walls.
Table 5.1 summarizes the major features for each side of the injector plate
evaluated by the new model at steady state conditions.
It is evident that the temperature at injector plate wall in the original version of

118
5. Transient Modelling

Figure 5.3.: Temperature profiles from original and new model

Table 5.1.: Injector plate variables comparison


Variable Fuel Oxidiser Chamber
Input
Propellant MMH N2 O4
Injector material Titanium Titanium
Injector head mass [kg] 1.5
Injector head thickness [m] 0.001 0.003
Injector head area [m2 ] 0.023 0.023 0.023
Chamber pressure [bar] 9.85
Chamber temperature [K] 3002.8
Mixture ratio [−] 1.65
2
hc coefficient [W/m · K] 450
Inlet temperature [K] 300 292.3
Output
Inj. heat flux [W] 17049 9835 26884
Cavities ∆T [K] 4.85 3.10
Wallc,hg temperature [K] 440.8 402.3 493.7

119
5. Transient Modelling

the model would have been unrealistic (Thg = Tc = 3002.8 K). Only the presence
of the virtual conductance allows the injector cavities not to increase the fluid
temperature to unrealistic values.
Using the newly developed model, the software is able to deliver reasonable
outputs with physically valid geometries. Moreover it is possible to obtain different
cavities wall temperatures for each propellant side, while before it could not occur.

5.2. Hot Gas side heat transfer coefficient models

5.2.1. Models implemented

In the ESPSS library the heat transfer coefficient inside the combustion chamber is
evaluated using the well-known Bartz correlation [9]. The original formulation of
this equation does not take into account several aspects, such as the combustion
zone due to atomization, vaporization and combustion delays in the proximity
of the injector plate, the boundary layer growth through the cylindrical part of
the chamber, the correct evaluation of the flow acceleration in the convergent-
divergent part of the nozzle, etc. . .
The heat flux in ESPSS takes into account the convective and radiative phenom-
ena:

4
q̇w = hc Awet (Taw − Tw ) + σAwet (Tcore − Tw4 ) (5.1)

Since many correction factors used in literature are based on Stanton type corre-
lations, it was decided to use this kind of dimensionless number to evaluate the
heat transfer coefficient.

hc q̇
St = = (5.2)
ρ∞ v∞ Cp,ref ρ∞ v∞ Cp,ref (Taw − Tw )

the Stanton number represents the ratio between heat transferred to a fluid and
the thermal capacity of this fluid. In the combustion chamber model three different
correlations have been implemented [35]:

• Original Bartz Equation

• Modified Bartz Equation

120
5. Transient Modelling

• Pavli Equation

In order to use the Bartz equation with Stanton type correction factors, the Bartz
equation has been rewritten as a Stanton type equation:

µ0.2
! 0.6 0.1
λref

ref −0.2 0.1 Dth π/4
StBartz = 0.026 0.6 (ṁ) A (5.3)
Cp,ref µref Rcurv

where the thermodynamic and transport properties are calculated at the so-called
film temperature calculated as: Tref = 0.5 (Tst + Tw ). To improve the behaviour of
the original Bartz equation, a temperature correction factor KT was added, taking
into account that the new reference temperature is calculated halfway between
the wall and the free stream static temperature. Moreover, since the geometric
reference parameter in the original Bartz equation was the throat diameter, a
further correction factor Kx was added for the consideration of the boundary
layer growth in the cylindrical part and in the nozzle [5]:
 a  b
Taw x
KT = Kx = (5.4)
Tref xth
hence, the Stanton type modified Bartz equation becomes:

StBartz,mod = StBartz KT Kx (5.5)

Because of its simplicity, the Pavli equation has been implemented as well. The
Pavli equation including the two correction factors discussed before is [103], [5]:
 e  f
−0.2 −0.6 Taw x
StP avli = 0.023 Re Pr (5.6)
Tref xth
The Reynolds number Re is calculated with respect to the local chamber di-
ameter and the property reference temperature is an averaged boundary layer
temperature. In this equation the temperature correction factor and the streamwise
correction factor are also included.
In order to improve the heat flux model in the combustion chamber another
correction factor was added taking into account the vaporization phenomenon
near the injector plate. In the so-called combustion zone the heat flux decreases
when getting closer to the injector plate. This behaviour is due to the incomplete

121
5. Transient Modelling

mixing and reaction of the flow for the given injector and combustion chamber.
The mixing region has a finite length where the combustion is less effective,
therefore the heat fluxes are lower. This correction factor is applied by using a
Stanton type correlation derived from Bartz or Pavli equations. The combustion
length can be given as an input or computed by a geometrical correlation. In
the latter case, to generate a specific correction factor two steps are required:
first, the length of this combustion zone xmax has to be calculated based on the
injector plate geometry; then a functional dependency of the heat flux in the range
x0 ≤ x ≤ xmax has to be found. Once the combustion zone length is evaluated
it is possible to calculate a correction factor by means of a tangential Stanton
number dependency [5]:

St∗ (x)
  
1 x
= arctan 7 − 0.63 + 0.7 (5.7)
St(xmax ) 4 xmax

The last correction factor added to achieve a better agreement between the
numerical and the experimental results is a correction factor Kacc related to the
flow acceleration. In fact, the measured heat fluxes are lower than the calculated
ones upstream and downstream the nozzle throat.
The behaviour is caused probably by the nozzle contour and therewith due to
the flow acceleration (bigger boundary layer thickness). Instead of using the local
velocity gradient to develop the correction factor, a more practicable way is to use
the absolute value of the first derivative of the chamber radius with respect to the
streamwise coordinate, |dr/dx|.
The correction factor Kacc requires two boundary conditions. In the cylindrical
part Kacc should be equal to 1. The other boundary condition is described by
Kacc = 0 and represents the disappearance of convective heat transfer due to flow
separation. The following correction is used to take into account both conditions
[5]: s

dr
St = St · Kacc = St · 1 − (5.8)
dx

122
5. Transient Modelling

5.2.2. Validation

In the years 1999-2000, in the frame of an ESA GSTP-2 contract, Astrium per-
formed a series of experiments with a water cooled calorimetric combustion
chamber [5]. The different correlations aforementioned based on Bartz ([5],[9])
and Pavli ([5],[103]) have been implemented to simulate the calorimetric combus-
tion chamber tests and their results have been compared with the experimental
results from this calorimetric chamber test campaign [57, 56].
The calorimetric chamber is a sub-scale, water cooled thrust chamber with
twenty segments [116]. Each segment features an independent water feed system
with volume flow measurement. For each segment the heat flux is measured
individually. The described calorimetric system has been modelled using the
following components [116]:

• 1 combustion chamber with 21 nodes (component to validate)

• 20 regenerative circuits with 5 nodes each

• 20 mass flow regulated water feed lines (with the necessary junctions and
boundary conditions)

• mass flow regulated propellant feed lines

The simulation was performed using the couple liquid oxygen/gaseous hydrogen
as propellants, at an O/F ratio varying from 5 to 7 and at a total pressure from 35
to 70 bar in the combustor; for each test point the propellant mass flows are chosen
in order to get the desired pressure and mixture ratio. In order to get the right
pressure drop through the cooling circuit, the roughness of the cooling channels
had to be adapted. Values between 3.2 and 25 µm were chosen. This tuning was
necessary because of the partially unknown layout of the cooling circuit and its
feed lines (pipes, fittings, . . . ).
The heat fluxes calculated with every correlation described in the previous
chapter are plotted in Figure 5.4 (a) for the nominal case (p = 60 bar, MR = 6);
in Figure 5.4 (b) simulation results are compared to experimental data obtained
varying the chamber pressure (p = 35, 60, 70 bar, MR = 6), Figure 5.4 (c) shows
the simulation and experimental comparison for tests with constant chamber

123
5. Transient Modelling

pressure but different mixture ratio (p = 60 bar, MR = 5, 7), while in Figure 5.4
(d) the hot gas wall temperature trend is shown at different chamber pressures.
The correlations are all plotted with lines and experimental data with symbols.

(a) Heat fluxes at MR = 6, pc = 60 bar (b) Heat fluxes at MR = 6, pc = 35, 60, 70 bar

(c) Heat fluxes at MR = 5, 7, pc = 60 bar (d) Wall temperatures at M R = 6, pc =


35, 60, 70 bar

Figure 5.4.: Heat fluxes and wall temperatures results

The “Combustion zone” correction factor is able to represent the lower heat
fluxes in the first part of the chamber. Unfortunately, the introduced correction
cannot be considered predictive (that is, a correction that would give good results

124
5. Transient Modelling

in a different combustion chamber and injector face): it would require experimen-


tal data with different calorimetric chambers and different injector configurations.
This is out of the scope of a 0-D/1-D investigation.
The heat fluxes in the divergent part are always overpredicted. This is a
characteristic of the Bartz model and needs to be kept in mind when interpreting
the results. However, introducing the “flow acceleration” correction factor is
possible to achieve a better agreement with the experimental results.
Moreover, unlike the combustion zone correction factor, its behaviour is not
peculiar of the experiment considered so it can be used for different chamber
configurations and performance conditions. No tuning has been performed on
the Bartz and Pavli parameters, the constants have been taken as C = 0.026, and
C = 0.023 respectively as recommended by Bartz and Pavli.
For each correlation, some remarks follow:

• Simple Bartz correlation. Here, the heat fluxes are underpredicted (around
30% in the cylindrical part) and the decreasing heat flux in the cylindrical
part is not shown, but the shape of the curve in the convergent divergent
nozzle region is similar to the experimental one.

• Modified Bartz correlation. Here, the heat fluxes are slightly overpre-
dicted, but using the temperature and the streamwise correction it has the
advantage of following very accurately the experimental data in the cylindri-
cal part. Therefore, the model without a “combustion zone” correction factor
can be applied only to part of the combustion chamber, after the mixing has
taken place.

• Pavli correlation. This correlation is able to follow the experimental trend


but in a different way of the Modified Bartz correlation. In fact, the Pavli
correlation underestimates the heat fluxes while the Bartz correlation over-
estimates them.

• Pressure dependency. Using the modified Bartz correlation, test cases at


different pressures have been modelled in EcosimPro. The results shown in
Figure 5.4 (b) indicate a very good agreement with experimental heat flux

125
5. Transient Modelling

values. Therefore, this correlation can be considered reliable for LOX/H2


combustion at MR = 6.

• Mixture ratio dependency. The same approach has been taken for the
mixture ratio dependency. Test cases at MR varying from 5 to 7 have been
modelled in the code. As can be seen in Figure 5.4 (c), the results present
a diverging behaviour. In particular, an increase in MR yields a general
increase in experimental heat fluxes, while the modified Bartz correlation
shows the opposite trend. It is difficult to indicate a clear explanation for
these results. The main drivers for the convective heat fluxes are the mixture
heat capacity at the reference temperature Cp,ref and the temperature gradi-
ent (Taw − Tw ). When MR increases, the heat capacity decreases (because of
less hydrogen in the mixture), whereas the temperature gradient increases.
In the modified Bartz model, it seems that of these two counteracting prop-
erties, the variation in Cp,ref is predominant. In the experiment, local MR
variations at the wall might be responsible for the opposed trend.

5.3. Q-2D stratification model for HARCC

For new engines the use of High Aspect Ratio Cooling Channels (HARCC) is
necessary. Indeed, the use of these kinds of channels permits a lower wall
temperature and a longer life. Beside these advantages, the HARCC have as
usual also drawbacks: the pressure drop is higher and thermal stratification
occurs within them. In order to optimize the design of this kind of channels it is
fundamental to evaluate the thermal stratification effect and so the heat absorption
of the coolant. It is therefore necessary to refine the models developed in the
system modelling tools in order to obtain more capabilities, using specific models
for each cooling system adopted.

5.3.1. Model description

As compared to two different papers from the Department of Mechanics and


Aerospace Engineering (DIMA) of “Sapienza” University of Rome [106] and the
German Aerospace Center (DLR) [150] that found their own way to analyse the

126
5. Transient Modelling

HARCC, a new approach [35] is here proposed to evaluate thermal stratification


in system tools such as EcosimPro. Starting from the one-dimensional governing
equations present in the ESPSS library:

∂u ∂f (u)
+ = S(u) (5.9)
∂t ∂x
where

   
ρ ρv
   
ρxnc   ρvxnc 
u = A ; f (u) = A  ; (5.10)
   
 ρv  2
ρv + P 
   
ρE ρvH
 
−ρAkwall (∂P/∂t)
 
 −ρx nc Ak (∂P/∂t) 
wall
S(u) =  (5.11)
 

−0.5(dξ/dx)ρ v|v|A + ρgA + P (dA/dx)
 
q̇w (dAwet /dx) + ρgvA

The new code presents an unsteady Q-2D model and can be considered as an
evolution of the two inspiring works presented by DIMA and DLR. The control
volumes are divided in slices, one on top of the other linked together longitudinally
by the momentum and energy viscous fluxes. The mass conservation equation
is written in a one-dimensional form but it is calculated for each slice, while the
momentum and energy conservation equations are written in a quasi-2D form
taking into account friction, longitudinal viscous transport, wall heat flux and
longitudinal fluid heat flux respectively.
Equations (5.9) and (5.10) have been modified in the following way, to obtain
inside each channel several longitudinal fluid veins one on top of the other and
linked by the momentum and energy viscous fluxes:

∂u ∂f (u) ∂g(u)
+ + = S(u) (5.12)
∂t ∂x ∂y
where

127
5. Transient Modelling

 
0
 
 0  ∂v ∂T
g(u) = Awet   ; τxy = µt ; qc = λt (5.13)
 
τxy  ∂y ∂y
 
qc

The turbulent conductivity coefficient λt is evaluated using the empirical cor-


relation of Kacynski [66]. By the use of a constant turbulent Prandtl number we
obtain the turbulent viscosity.

λt P rt λt
= 0.008 Re0.9 P rt = 0.9 µt = (5.14)
λ cp
Hence each slice has his own velocity, and no empirical correlations are used
to evaluate the velocity profile being automatically related to the viscous fluxes
and the longitudinal heat flux. To accurately describe the wall heat flux also the
wall temperature is assumed to vary along the y direction. All thermodynamic
properties such as temperature, density and enthalpy depend on x, y and time.

(a) 1D Fluid Element (b) Q-2D Fluid Element integrated


with walls

Figure 5.5.: left: 1-D fluid element and energy balance used for conventional 1-D
method; right: control volumes of the Q-2D approach integrated in 3D
wall elements

128
5. Transient Modelling

Initial and boundary conditions of the cooling channel are the typical ones for
capacitive components: a capacitive component receives the flow variables (mass
and enthalpy flows) as input in inlet and outlet and gives back the state variables
(pressure and enthalpy) as output.
The cooling channel model is built from 3 components: one quasi-2D tube and
two volumes, one at the inlet and the other one at the outlet, representing the
manifold volumes of a typical cooling jacket. Each slice is connected directly to
the volumes. The quasi-2D tube is a resistive component: it receives the state
variables as input in inlet and outlet and gives back the flow variables as output.
Please note that no velocity profile in the y direction has been assigned, but
each fluid vein is affected by viscous fluxes and wall friction. Moreover a real
time-dependent integration has been performed, in order to evaluate the thermal
stratification through the time for unsteady analysis.
The evaluation of the friction factor of each cell has been done by using a
peculiar hydraulic diameter defined as function of the wet channel surface and
the perimeter of each volume:

4Ai
Dh,i =
Pwet,i

To our knowledge it is the first time that a quasi-2D approach is implemented for
pipe flows in a system tool for transient analysis; with this model we are able to
evaluate not only the stratification effect but also the time that the coolant needs
to show this stratification during the transient phase of the engine ignition.

3-D cooling channels walls

The “3D wall” components used to simulate the walls are part of the original
ESPSS library [42]. They will calculate the heat conduction in every direction
including the axial direction. This thermal component features thermal ports in
radial and in azimuth directions allowing an exact calculation of heat conduction
through the channel corners. The model has been modified in order to allow the
connection between its thermal ports and the quasi-2D channel ports. The walls
are divided in 5 different 3-D components as shown in Figure 5.6. Each component
has a 3-dimensional discretisation in tangential, radial and longitudinal direction

129
5. Transient Modelling

Figure 5.6.: Cooling jacket wall mesh

(dx, dy , dz ), respectively.
The formulation for this component is the typical one for conduction elements;
the thermal capacitance for each volume is defined as:

Ci,j,k = ρ Cp(i,j,k) dx dy dz (5.15)

the internal heat flows are evaluated by:

q̇x(i,j,k) = ki,j,k dy dz (Ti−1,j,k − Ti,j,k )/dx (5.16)


q̇y(i,j,k) = ki,j,k dx dz (Ti,j−1,k − Ti,j,k )/dy (5.17)
q̇z(i,j,k) = ki,j,k dx dy (Ti,j,k−1 − Ti,j,k )/dz (5.18)

while the energy equation is:

dTi,j,k
Ci,j,k = q̇x(i,j,k) − q̇x(i+1,j,k) + q̇y(i,j,k) − q̇y(i,j+1,k) + q̇z(i,j,k) − q̇z(i,j,k+1) (5.19)
dt

As shown in Figure 5.6 only half channel has been considered because of
symmetry reasons, with left and right sides adiabatic:

130
5. Transient Modelling

q̇out,right_r = 0 (5.20)
q̇out,int_l = 0 q̇out,int_right_r = 0 (5.21)
q̇out,ext_l = 0 q̇out,ext_right_r = 0 (5.22)

5.3.2. Numerical validation

The Q-2D model for cooling channels has been validated by comparison with a
numerical test case performed by DIMA [107, 108] of a turbulent flow of methane
in a straight channel with asymmetric heating. These calculations have been
compared with ESPSS 1D calculations and with the new Q-2D model object of this
validation. The channel is smaller than the ones used in actual rocket channels.
Indeed, the geometric and the boundary conditions have been chosen by DIMA to
obtain small values of the Reynolds number, because the computational grid size
of the 3D CFD code is function of this parameter [109].
In order to validate the correct behaviour of the new transient model, two
different aspect ratios of the channel have been investigated, a first channel with
aspect ratio 1 and a second one with aspect ratio 8. The length and the cross
section area of the channel have been kept the same among the two different
channels. Both channels are 27 mm long and have a cross section of 0.08 mm2 .
The boundary conditions are the same for both channels and for all models:
a stagnation inlet temperature of 220 K, a stagnation inlet pressure of 90 bar, a
constant temperature of 600 and 220 K at the bottom and at the top of the walls,
respectively. Along the lateral side of the channel a linear temperature distribution
is applied from 600 to 220 K.
At the inlet of the channel a pressure source provided the inlet pressure, while
at the outlet a mass flow controlled component forced the mass flow rate. The
outlet pressure is an output of the model. Three different temperature sources
provided the correct temperature distribution for the bottom side, the lateral side
and top side of the channel respectively. To ensure a correct trend of thermal
sources during the transient phase, a conductive and a capacitive component have
been linked between each temperature source and the thermal ports of the channel.
The same configuration has been applied for the aspect ratio 1 and the aspect ratio

131
5. Transient Modelling

8 channel.

Results

Figures 5.7 (c,d) show the bulk evolution of the pressure and temperature along
the channel for the aspect ratio 1 case, while Figures 5.7 (e,f) show the pressure
and temperature evolution for the aspect ratio 8 channel.
When the stratification effect is not so evident, as in the aspect ratio 1 case, the
1D model and Q-2D model have a similar trend; but when stratification occurs,
as in the aspect ratio 8 channel, the differences among 1D and Q-2D model are
evident, and the Q-2D results are closer to the 3D-CFD ones.
Figure 5.7 (a,b) compares the cross-section temperature contours at the channel
outlet, for each studied model and for both aspect ratios discussed here. The
AR = 1 case features some temperature stratification in the 3D simulation. This
has not been observed with the Q-2D model described in Section 5.3, which shows
virtually no stratification. On the other hand, for AR = 8, where a consistent
stratification is expected, a very good agreement can be observed between the 3D
simulations and the new Q-2D model.

5.3.3. Experimental validation

The DLR Lampoldshausen test bench features a cylindrical combustion chamber


segment with four different cooling channel geometries used to investigate thermal
stratification [151, 132]. Its test results have been used to validate our Q-2D model
for high aspect ratio cooling channels. The combustion chamber was designed at
DLR institute of Space Propulsion particularly for studies with interchangeable
segments.
The combustor has a combustion chamber internal diameter of 80 mm and a
nozzle throat diameter of 50 mm. Liquid hydrogen is supplied to the combustor at
temperatures as low as 50-60 K while supply pressures are in the range of 200-250
bar. The HARCC segment is a single cylindrical segment with a diameter of 80
mm and 209 mm length. The test segment has on its circumference four different
cooling channel geometries, in each 90◦ sector the cooling ducts have a different
aspect ratio.

132
5. Transient Modelling

(a) Temperature stratification, AR = 1 (b) Temperature stratification, AR = 8

(c) Bulk temperature, AR = 1 (d) Bulk pressure, AR = 1

(e) Bulk temperature, AR = 8 (f) Bulk pressure, AR = 8

Figure 5.7.: Methane bulk variables evolution along channel axis

133
5. Transient Modelling

Figure 5.8.: Design of the 4 sector HARCC segment

section height [mm] width [mm] channels number AR


3 9.0 0.3 152 30
4 4.6 0.5 136 9.2

Table 5.2.: Cooling channels geometries

Figure 5.8 shows the construction of the HARCC-segment with different cooling
channel geometries. The experiment was performed using the couple liquid
oxygen/gaseous hydrogen as propellants, and liquid hydrogen as coolant.
Two pressure configurations have been simulated: the first with a chamber
pressure of about 88 bar, and the second with a chamber pressure of 58 bar. Two
sectors have been investigated: Quadrant 4, with channel aspect ratio 9.2 and
Quadrant 3 with channel aspect ratio 30.
The geometry of the investigated cooling channels as well as the number of
channels referred to circumference of the chamber are given in Table 5.2.
For each Quadrant, four sets of thermocouples have been positioned along
the channels. In each group, 5 thermocouples have been arranged with different

134
5. Transient Modelling

Position 1 2 3 4
Distance from leading edge of the segment [mm] 52 85 119 152
Thermocouple TE1 TE2 TE3 TE4 TE5
Distance from the hot gas wall [mm] 0.7 1.1 1.5 1.9 7.5

Table 5.3.: Positioning of themocouples

distances from the hot gas side wall. Location and distance from the wall of the
thermocouples are summarised in Table 5.3.
Such temperature measurements at different locations provide important infor-
mation regarding the development of stratification along the channels.

Modelling

The test bench has been modelled using EcosimPro [36]. As shown in Figure 5.9,
two mass flow sources provide the correct mass flow rate of oxygen and hydrogen
to the combustion chamber component. Because the HARCC test segment repre-
sents only a portion of the cylindrical part of the combustor, the first segment has
been modelled as adiabatic.
Thermal demux components connect the combustion chamber to the HARCC
segment. At the inlet of the channel a pressure source provided the inlet pressure,
while at the outlet a mass flow source forced the mass flow rate. The channel
walls features three different materials: the inner side and the fins are in copper
alloy; the external wall is built with another copper alloy and a jacket in Nickel
alloy.

Results

Figures 5.10 and 5.11 refer to Quadrant 4 with channel aspect ratio of 9.2 and
show the simulation results for the pc = 88 bar test and the pc = 58 bar test,
respectively.
Figures 5.12 and 5.13 refer to Quadrant 3 with channel aspect ratio of 30 instead
and show the simulation results for the pc = 88 bar test and the pc = 58 bar test,
respectively.
Figures 5.10 (a,b) compare the wall temperatures in the cooling channels and

135
5. Transient Modelling

Figure 5.9.: Schematic of the experimental test case

136
5. Transient Modelling

the fluid temperatures, respectively, obtained by 1D and Q-2D simulations for the
high pressure test case, while Figures 5.10 (c,d,e,f) show the temperature values
at thermocouples positions, comparing experimental values with Q-2D and 1D
simulation results. Figure 5.11 shows the same variables for the low pressure test
case.
Hydrogen enters the channels in supercritical conditions. The hot gas side heat
transfer correlation described by Eq. (6.14) was slightly adapted to the calculated
experimental hot gas side heat fluxes. In Equation 6.14, an adapted value of 0.0263
was taken. Hence representative hot gas conditions have been modelled in terms
of heat transfer coefficient and combustion chamber temperatures.
From Figures 5.10 (a,b) it is evident that the behaviour of the 1D model is
completely different from the Q-2D model. The Q-2D model is able to obtain
a more representative temperature trend in the radial and in the longitudinal
direction. From the contour plot it is clear that the 1D model provides a very
homogeneous temperature profile also in the walls because it is not able to take
into account the occurring of thermal stratification. The validity and the usefulness
of the Q-2D model is enhanced by the comparisons shown in Figures 5.10 (c,d,e,f):
when high aspect ratio is used, 1D models are not adequate any more.
Figures 5.12 and 5.13 show the same variables for channel aspect ratio 30. In
these figures the difference between the Q-2D and 1D behaviour compared to
experimental data is once more evident.
Looking at Figures 5.12 (c,d,e,f) the maximum percentage error obtained by the
Q-2D model, when compared to the experimental data of the first thermocouple,
does not exceed 10%, while the percentage error for the 1D model is around 40%.
Better results we obtain if we compare the percentage error of the same variable
in the 58 bar test case, where Q-2D model error does not exceed 5% and 1D model
error is around 39%.

137
5. Transient Modelling

(a) Wall temperatures, AR = 9.2, pc = 88 bar (b) Fluid temperatures, AR = 9.2, pc = 88 bar

(c) Thermocouples temperatures, x = 52 mm, (d) Thermocouples temperatures, x = 85 mm,


pc = 88 bar pc = 88 bar

(e) Thermocouples temperatures, x = 119 mm, (f) Thermocouples temperatures, x = 152 mm,
pc = 88 bar pc = 88 bar

Figure 5.10.: Wall and fluid thermal stratification,AR = 9.2, pc = 88 bar

138
5. Transient Modelling

(a) Wall temperatures, AR = 9.2, pc = 58 bar (b) Fluid temperatures, AR = 9.2, pc = 58 bar

(c) Thermocouples temperatures, (d) Thermocouples temperatures, x = 85 mm,


x = 52 mm,pc = 58 bar pc = 58 bar

(e) Thermocouples temperatures, (f) Thermocouples temperatures, x = 152 mm,


x = 119 mm,pc = 58 bar pc = 58 bar

Figure 5.11.: Wall and fluid thermal stratification, AR = 9.2, pc = 58 bar

139
5. Transient Modelling

(a) Wall temperatures, AR = 30, pc = 88 bar (b) Fluid temperatures, AR = 30, pc = 88 bar

(c) Thermocouples temperatures, (d) Thermocouples temperatures, x = 85 mm,


x = 52 mm,pc = 88 bar pc = 88 bar

(e) Thermocouples temperatures, (f) Thermocouples temperatures, x = 152 mm,


x = 119 mm,pc = 88 bar pc = 88 bar

Figure 5.12.: Wall and fluid thermal stratification, AR = 30, pc = 88 bar

140
5. Transient Modelling

(a) Wall temperatures, AR = 30, pc = 58 bar (b) Fluid temperatures, AR = 30, pc = 58 bar

(c) Thermocouples temperatures, (d) Thermocouples temperatures, x = 85.8 mm,


x = 52.5 mm,pc = 58 bar pc = 58 bar

(e) Thermocouples temperatures, (f) Thermocouples temperatures,


x = 119.1 mm,pc = 58 bar x = 152.5 mm, pc = 58 bar

Figure 5.13.: Wall and fluid thermal stratification, AR = 30, pc = 58 bar

141
6. Integrated Validation: RL-10 design and
analysis

The RL-10 engine is based on an expander cycle, in which the fuel (H2 ) is used
to cool the main combustion chamber and the thermal energy added to the fuel
drives the turbopumps. The RL-10 rocket engine is an important component of
the American space infrastructure. Two RL-10 engines form the main propulsion
system for the Centaur upper stage vehicle, which boosts commercial, scientific
and military payloads from a high altitude into Earth orbit. The RL-10A-3-3A
developed by Pratt & Whitney under contract to NASA, incorporates component
improvements with respect to the initial RL-10A-1 engine.
A cryogenic expander cycle engine involves a strong coupling between the dif-
ferent subsystems. This coupling is even stronger during the start and shut-down
transients, when non-linear interactions between subsystems play a major role.
In addition complex phenomena such as combustion, heat transfer, turbopump
operation, phase change, valve maneuverings are concerned, as well as important
changes in the thermodynamic properties of the fluids involved. A transient
model helps to reduce the number of engine tests by allowing to perform a certain
amount of parametric studies in advance of the test campaign, and thus plays an
important role in the cost and risk reduction.
The RL-10 engine has been used extensively as object of simulations in the past
years [15, 14, 13, 59, 58]. In this chapter we want to show the improvements made
in terms of modelling with respect to the other tools; indeed, the model presented
here features a 1-D discretisation not only in the cooling jacket model, but also
for most of the other components, such as the combustion chamber, the Venturi
duct and the other pipes.
In previous works [15], the combustion chamber has been modelled as a built-in
set of hydrogen/oxygen combustion tables. Here, a fully 1-D discretised chamber

142
6. Integrated Validation: RL-10 design and analysis

and nozzle features a chemical equilibrium model based on Gibbs energy mini-
mization for each section along the chamber. The present model also contains the
injector plate model described in chapter 5.1 representative not only of the capaci-
tive effect of the injector dome mass but also of the convective and radiative heat
fluxes from the chamber to the injector and of the conductive heat flux between
the fuel and oxidiser injector domes. The thermal model used for the cooling
jacket component is modelled as a “real” one and a half counterflow cooling jacket.
Finally it is important to mention that, to the best of the author’s knowledge,
chill down and pre-start procedures were never simulated before with transient
system tools for the whole engine. In the present work, the cool-down (pre-
start) procedure has been simulated in order to obtain a accurate and complete
engine state at start signal (t=0). The pre-start simulation results are in very good
agreement with the few experimental data available.

6.1. Overview of the RL-10A-3-3A rocket engine

The RL10A-3-3A includes seven engine valves as shown in Figure 6.2. The
propellant flows to the engine can be shut off using the Fuel Inlet Valve (FINV)
and the Oxidizer Inlet Valve (OINV). The fuel flow into the combustion chamber
can be stopped by the Fuel Shut-off Valve (FSOV) located just upstream of the
injector plenum. The FSOV is a helium operated, two position, normally closed,
bullet-type annular gate valve. The valve serves to prevent fuel flow into the
combustion chamber during the cool-down period and provide a rapid cut-off of
fuel flow during engine shut-down [110].
The fuel interstage and discharge cool-down valves (FCV-1 and FCV-2) are
pressure-operated, normally open sleeve valves. The purposes of these valves are
the following [110]:

• allow overboard venting of the coolant for fuel pump cool-down during
engine pre-chill and pre-start

• provide first stage fuel pump bleed control during the engine start transient
(for the FCV-1)

• provide fuel system pressure relief during engine shut-down

143
6. Integrated Validation: RL-10 design and analysis

The Thrust Control Valve (TCV) is used to control thrust overshoot at start
and maintain constant chamber pressure during steady-state operation. TCV is a
normally closed, servo-operated, closed-loop, variable position bypass valve used
to control engine thrust by regulation of turbine power. As combustion chamber
pressure deviates from the desired value, action of the control allows the turbine
bypass valve to vary the fuel flow through the turbine [110].
The Oxidizer Control Valve (OCV) has two orifices: one regulates the main
oxidizer flow (OCV-1) and the other controls the bleed flow required during engine
start (OCV-2). The main-flow orifice in the OCV is actuated by the differential
pressure across the LOX pump. The OCV valve is a normally closed, variable
position valve. The valve controls oxidiser pump cool-down flow during the
engine pre-start cycle and during engine start transient [110].
The Venturi upstream of the turbine is designed to help stabilize the thrust
control.
Ducts and manifolds in the RL10 are generally made out of stainless steel and
are not insulated.
The combustion chamber and nozzle walls are composed of cooling tubes. A
silver throat is cast in place for the RL10A-3-3A and increases the expansion ratio
for higher specific impulse. The inject has 216 coaxial elements; the oxidiser is
located in the center of each element and hydrogen through the annulus. One-
hundred-sixty-two of the LOX injector elements have ribbon flow-swirlers that
provide enhanced combustion stability.
The regenerative cooling jacket serves several functions in the RL10 engine.
The basic configuration is a pass-and-a-half stainless-steel tubular design. Fuel
enters the jacket via a manifold located just below the nozzle throat. A set of 180
“short” tubes carry coolant to the end of the nozzle. At the nozzle exit plane, a
turn-around manifolds directs the flow back through a set of 180 “long” tubes. The
long tubes are interspersed with the short tube in the nozzle section and comprise
the chamber cooling jacket above the inlet manifold. Coolant flow exits through a
manifold at the top of the chamber. The cooling tubes are brazed together and act
as the inner wall of the combustion chamber and nozzle.
The fuel pump consists of two stages, separated by an interstage duct, which is
vented via the interstage cool-down valve (FCV-1) during start. Both fuel pump

144
6. Integrated Validation: RL-10 design and analysis

stages have centrifugal impellers, vaneless diffusers and conical exit volutes; the
first stage also has an inducer.
The LOX pump consists of an inducer and a single centrifugal impeller, followed
by a vaneless diffuser and conical exit volute. The LOX pump is driven by the
fuel turbine through the gear train. The turbopump speed sensor is located on the
LOX pump shaft [111].
The RL-10 turbine is a two stage axial-flow, partial admission, impulse turbine.
Downstream of the turbine blade rows, exit guide vanes reduce swirling of the
discharged fluid. The turbine is driven by hydrogen and powers both fuel and
oxidiser pumps.
There are a number of shaft seals which permit leakage from the pump discharge
in order to cool the bearings. The fuel pump and the turbine are on a common
shaft; power is transferred to the LOX pump through a series of gears. The seals,
bearings, gear train all contribute to rotordynamic drag on the turbopump.

Figure 6.1.: RL-10A-3-3A engine schematic [115]

145
6. Integrated Validation: RL-10 design and analysis

Name Value Units


Fuel Turbopump
1st stage impeller diameter 179.6 [mm]
1st stage exit blade height 5.8 [mm]
2nd stage impeller diameter 179.6 [mm]
2nd stage exit blade height 5.588 [mm]
Oxidiser Turbopump
Impeller diameter 106.7 [mm]
Exit blade height 6.376 [mm]
Turbine
Mean line diameter 149.86 [mm]
Mass moment of inertia 0.008767 [kg·m2 ]
Ducts & Valves
FINV flow area 0.0041 [m2 ]
FCV-1 flow area 0.00038 [m2 ]
FCV-2 flow area 0.00019 [m2 ]
Pump discharge duct 0.0011 [m2 ]
Venturi (inlet - throat) 0.0023 - 0.00067 [m2 ]
TCV flow area a 1.01E−5 b [m2 ]
Turbine discharge housing (inlet - exit) 0.013 - 0.003 [m2 ]
Turbine discharge duct 0.003 [m2 ]
FSOV flow area 0.0021 [m2 ]
OINV flow area 0.0031 [m2 ]
OCV flow area a 3.96E−4 b [m2 ]
Cooling jacket
Number of short tubes 180 [-]
Number of long tubes 180 [-]
Channel width at throat 2.286 [mm]
Channel height at throat 3.556 [mm]
Total coolant volume 0.0158 [m3 ]
Typical hot wall thickness 0.3302 [mm]
HGS effective surface area 4.645 [m2 ]
Thrust chamber
Chamber diameter 0.1303 [m]
Throat diameter 0.0627 [m]
Nozzle area ratio 61 [-]
Chamber/nozzle length 1.476 [m]
Number of injectors 216 [-]
Injector assembly weight 6.72 [kg]

Table 6.1.: RL-10A-3-3A construction data [15]


a values at nominal full-thrust condition
b
this flow area includes the discharge coefficient for the orifice, which is unknown

146
6. Integrated Validation: RL-10 design and analysis

Figure 6.2.: RL-10A-3-3A engine diagram

6.2. Design procedure

The development of the RL-10 engine transient model has been conducted with
EcosimPro and the ESPSS library, in the upgraded version including all the
relevant models developed and described in Chapter 5.

6.2.1. Turbomachinery modelling

Pumps
The pump model makes use of performance maps for head and resistive torque.
The pump curves are introduced by means of fixed 1-D data tables defined as
functions of a dimensionless variable θ that preserves homologous relationships
in all zones of operation. θ parameter is defined as follows:

θ = π + arctan(ν/n) (6.1)

147
6. Integrated Validation: RL-10 design and analysis

where ν and n are the reduced flow and the reduced speed respectively:

Q ṁin /ρin 30 ω /π
ν= = n= (6.2)
QR QR rpmR
The dimensionless characteristics (head and torque) are defined as follows:

TDH / TDHR τ / τR
h= β= (6.3)
n2 + ν 2 n2 + ν 2
this method eliminates most concerns of zero quantities producing singularities.
To simplify the comparison with generic map curves, these relations are normal-
ized using the head, torque, speed and volumetric flow at the point of maximum
pump efficiency. These maps have been created as a combination of available test
data provided by Pratt&Whitney [15] and generic pump performance curves [25]
(see Figures 6.3, test data range in grey). Additional maps were established (not
shown here), giving a corrective factor on the pump torque, function of the rota-
tional speed ratio (also provided by P&W). The enthalpy flow rise is a function
of the absorbed power while the evaluation of the mass flow rate is performed
through an ODE.

(ṁ h)out = τ · ω − (ṁ h)in

   
dṁ 1 2 1 2
I ·
 = P + ρv − P + ρv − gρin · TDH
dt 2 out 2 in

Because of the presence of the FCV-2 valve between the first and the second
stage, the fuel pump has been modelled with two separated pump components, one
for each stage. Since the oxidiser pump has only one stage, it has been modelled
with one component instead. For each pump model the main nominal parameters
have been calculated by a numerical code specifically developed to find the
nominal value of the outlet pressure, the pump torque τp , the total dynamic head
TDH , the pump efficiency ηp and the specific speed Ns by use of the Pump head
and Pump efficiency curves provided by Pratt & Whitney [15]. The development
of a dedicated tool for the evaluation of the pump nominal parameters has been
necessary since there was a discrepancy between the definition of total dynamic

148
6. Integrated Validation: RL-10 design and analysis

(a) Extended Head map for LOX and Fuel (b) Extended Torque map for LOX and Fuel
pumps pumps

Figure 6.3.: Pumps performance maps

head used in the pump model and the one used in P&W maps:

hout − hin Pout − Pin


TDHtool = ; TDHESP SS = ; (6.4)
g ρin g
 
Pout Pin .
TDHP &W = − g (6.5)
ρout ρin

The first one represents the head rise given by the enthalpy difference between
the inlet and the outlet conditions; this definition has been used to match the
requested power of the pump. The second one is the head given by the pressure
difference between inlet and outlet and the inlet density; this is the definition
used in the ESPSS pump model (see Eq. 3.32). The third one is defined using the
difference between the pressure on density ratio at outlet and inlet and used to
define the numerical value from the P&W maps.
These three definitions of the dynamic head can be considered the same only in
the ideal case of a pure incompressible fluid (ρin = ρout ).
As real fluids in the pump component are used, even if the fluid is in liquid
conditions, the density difference between inlet and outlet generates a discrepancy
between the aforementioned definitions. Moreover, using in the tool the Euler
equation of turbomachinery to calculate the power, and comparing it with Eq. 3.30

149
6. Integrated Validation: RL-10 design and analysis

we obtain:

∆his (hout − hin )is


W = ṁ∆h = ṁ =τω ⇒ QR ρin g · = τR · ω (6.6)
ηp η

(hout − hin )is TDHR


The term = ⇒ TDH is not the same of the one
η η
present in “inlet mass flow equation” (see Eq. 3.32); in fact, comparing the defini-
tion of TDH in Eq. 3.30 with the definition of TDH in Eq. 3.32 we obtain:

∆P
TDHR |Eq. 3.32 = 6= (hout − hin ) = TDHR |Eq. 3.30 (6.7)
ρg

The mismatch present in the use of two different versions of the total dynamic
head could affect the results of the simulations and the performances of pump
itself. For this reason the code developed is able to calculate a “modified” pump
efficiency in order to match either the pressure rise either the pump torque in the
ESPSS pump model. Since no official values of the propellants leak to the gear box
were found, an iterative procedure was adopted to find the correct value of the
mass flow rate and the outlet pressure in each stage.

∆h = W/ṁ hout
NO

Pout , ṁ
Input Pout ? YES
hout,is = TDHcalc = TDHP &W hout , η
W , ṁ0 , Pin , Tin f (hout,is , sin )
TDH

TDHP &W NO
φ = f (ṁ) ηP &W

Figure 6.4.: Iterative procedure for determining pump parameters

Turbine
The turbine performance maps provided by Pratt&Whitney depict the combined
performance of the two stages (see Figures 6.5 (a,b)). The first one describes
the effective area (area times discharge coefficient) as a function of velocity

150
6. Integrated Validation: RL-10 design and analysis

ratio (U/Co ) for several different pressure ratios. The second one describes the
combined two-stage turbine efficiency as a function of velocity ratio (U/Co ) as
well.

(a) Efficiency map for the Turbine (b) Effective Area map for the Turbine

Figure 6.5.: Turbine performance maps from P&W [15]

In the present study, Pratt & Whitney performance maps are transformed to
obtain the turbine performance maps used in the ESPSS turbine model. These
maps (mass flow coefficient and specific torque) are introduced by means of 2-D
input data tables as a function of velocity ratio and pressure ratio (see Figures 6.6
(a,b)):

r·ω
N= Π = P01 / P02 (6.8)
Co
and the mass flow coefficient and specific torque are defined as:

ṁmap · Co τ
Q+ = ST = (6.9)
r2 P01 r ṁmap Co
According to Eq. 6.9 and to the power balance equation τ · ω = ṁ η ∆his we
obtain the non-dimensional parameters as function of velocity ratio and pressure

151
6. Integrated Validation: RL-10 design and analysis

ratio using data from the P&W maps:

τ · ω = ṁ η(Π) ∆his ⇒ ST r ṁ Co · ω = ṁ η(Π) ∆his


⇒ ST Co2 N = η(Π) ∆his ⇒
η(Π) ∆his
ST (Π, N ) =
Co2 N

and for the Q+ parameter we just need to calculate the turbine mass flow as
function of N and Π:

Aef f = CD · A = f (Π, N )
r
2γ h − γ2 γ+1
i
ṁ = CD · A · P01 ρ01 Π − Π− γ
γ−1

This formulation is based on the assumption that no chocking conditions occur


during the transient and at steady conditions of the turbine component.

(a) Specific Torque map for the Turbine (b) Mass Flow coefficient map for the turbine

Figure 6.6.: Turbine performance maps

6.2.2. Thrust chamber and cooling jacket modelling

The thrust chamber component, inherited from the original ESPSS library [43],
represents a non adiabatic 1-D combustion process inside a chamber for liquid

152
6. Integrated Validation: RL-10 design and analysis

or gas propellants. The equilibrium combustion gases properties (molar fraction,


thermodynamic and transport properties) are calculated for each chamber volume
(node) using the minimum Gibbs energy method [55] as a function of the propel-
lant’s mixture molar fractions, inlet conditions and chamber pressure. Transient
chamber conditions (pressures, temperatures, mass flows and heat exchanged with
the walls) are derived from 1-D transient conservation equations (refer to sec-
tion 3.4). A mixture equation between the injected propellants and the combustion
gases is applied. From the definition of the mixture ratio MR and derivation, the
following dynamic equation gives the MR evolution:

d ρVc
ṁox = M R ṁf u + (M R) (6.10)
dt 1 + MR
Combustion takes place when mixture ratio is within the allowed limits, the
ignition flag is active and a minimum time (ignition delay) τ has elapsed. Mass,
energy and momentum equations are basically the same as in the pipe component
with variable cross area, Equations (6.11), (6.12).

∂u ∂f (u)
+ = S(u) (6.11)
∂t ∂x
where

   
ρ ρv
   
ρxnc   ρvxnc 
u = A ; f (u) = A  ;
   
 ρv  ρv 2 + P 
   
ρE ρvH
 
0
 
 0 
S(u) =  (6.12)
 

−0.5(dξ/dx)ρ v|v|A + ρgA + P (dA/dx)
 
q̇w (dAwet /dx) + ρgvA

The centred scheme is used to discretise the chamber, using a staggered mesh
approach (see Figure 6.7). The chamber contour has been divided in 40 volumes:
10 in the subsonic section, 10 from the throat to cooling jacket inlet manifold

153
6. Integrated Validation: RL-10 design and analysis

and the last 20 volumes from there until the nozzle exit. The mesh has been
stretched and compressed in order to capture the main fluid-dynamic phenomena
occurring along the chamber (fluid acceleration, heat flux in the throat region,
Mach evolution). The RL10A-3-3A has a silver throat insert that creates a sharp

Figure 6.7.: RL-10A-3-3A chamber contour [15] and discretisation

edge, not typically used and difficult for EcosimPro to model. For this reason a
scale coefficient factor named Rins has been added into the code; the coefficient is
function of the silver insert geometry and the effective throat area considering the
reduction due to viscous effects.
The walls represented by thermal components in the Cooling Jacket component
are not included in the chamber model, but are taken as a boundary for the heat
exchange calculation instead:

4
q̇w = hc Awet (Taw − Tw ) + σAwet (Tcore − Tw4 ) (6.13)

154
6. Integrated Validation: RL-10 design and analysis

In the combustion chamber component the heat transfer coefficient hc can


be evaluated by different correlations (original Bartz equation, modified Bartz
equation, Pavli equation). Refer to chapter 5.2 for a detailed description of the heat
transfer correlation models. An heat transfer simulation campaign at subsystem
level has been performed in order to compare the different correlations and choose
the most suitable. Then the modified Bartz equation has been chosen. The Bartz
equation has been rewritten in a Stanton type form and modified with correction
factors:

µ0.2
! 0.6 0.1
λref

ref −0.2 0.1 πDth /4
StBartz = 0.026 (ṁ) A KT Kx (6.14)
c0.6
p,ref µref Rcurv

The RL10A-3-3A injector plate is rather complicated, involving several different


injector element designs. Most of the injector elements are co-axial, the hydrogen
in injected through annular orifices around each LOX element. The outer con-
centric row of elements, however, inject hydrogen only (which will affect wall
cooling). It is possible that some of the differences encountered in the heat transfer
model (see section 5.1) are due to not including this film cooling effect in those
predictions.
The injector plate composed by injectors and injector domes is modelled by
a component that takes into account the convective and radiative heat transfer
between the fluid in the first volume of the chamber and the face plate, and
evaluates the conductive and capacitive effect of the injector walls in an accurate
way, representative of a generic injector head (refer to section 5.1). In order to
reflect the thermal capacity of the injector plate, the actual mass and the material
properties of the dome have been used into the model (see Table 6.1). For the
oxidiser and fuel injector orifices, junctions components have been specifically
modelled to match the mass flow and the pressure drop. Nevertheless, the geomet-
rical construction data of the injector orifices have not been modified but used to
assess the pressure drop coefficient ζ ; for each propellant injectors, considering
the orifice area as the sum of the overall injectors, it yields:

ṁ2
   
Pcc 1 2 Pcav 1
+ v − = − ζv 2 ⇒ ∆P = (1 + ζ)
ρ 2 ρ 2 2ρA2

155
6. Integrated Validation: RL-10 design and analysis

The cooling jacket model is constructed of 360 stainless steel tubes of type 347SS
properties. There are 180 short tubes, from inlet manifold to the turn-around one,
and other 180 long tubes, from the turn-around manifold to the injector plate. The
short and long tubes are arranged side-by-side in the nozzle section.
A new model structure has been developed and implemented just for the RL-10
cooling jacket subsystem. The model has been built with two Tube components,
the first one simulating the short channels and the second describing the long
channels. The two tubes are connected together thanks to a Junction component
that models in this way the pressure drop caused by the turn-around manifold.
The component developed is able to reproduce the peculiar pattern of the cooling
channels in the nozzle section, where the long tubes are interspersed with the short
tubes. The heat coming from the chamber is then distributed to both channels
respectively.
The cooling jacket model is divided into a variable number of sections in axial
direction. Every section is made of one fluid node of the Tube component (from
FLUID_FLOW_1D library, see Equations (6.11),(6.12)), which is simulating the
cooling channels and five slices of the “3D wall” components, which are simulating
the metallic walls. The walls are divided in 5 different 3-D components as shown
in Figure 3.5; the contours of the actual height and width of the RL-10 channels
are shown in Figure 6.8 (a, b). Each component has a 3-dimensional discretisation
in tangential, radial and longitudinal direction (dx, dy , dz ), respectively.
Since the cooling channel shape is not rectangular but slightly rounded (see
Figure 6.8 (c)), a detailed geometrical reconstruction has been performed to assess
the effective exposed surface area, to maintain the original pressure drop and the
coolant velocity evolution. To this purpose the Pratt&Whitney specification has
been accepted regarding the angle of exposure which is around 112◦ [15].

156
6. Integrated Validation: RL-10 design and analysis

(a) Short channels width and height pro- (b) Long channels width and height profile [137]
file [137]

(c) Detail of Tubular Construction

Figure 6.8.: Cooling jacket channels profiles

157
6. Integrated Validation: RL-10 design and analysis

6.2.3. Lines, valves and manifolds modelling

In addition to the various subsystem listed above, there are on the RL-10 engine
a large number of lines valves and manifolds. Valves are modelled as zero di-
mensional components while the lines present in the engine are modelled via
an area-varying non-uniform mesh 1-D scheme. Where possible and data were
available a detailed geometrical reconstruction has been performed, as for the case
of the Venturi pipe and the discharge turbine pipe.

Fuel line set


The fuel line setup enables the fuel flow from the FINV to the combustion
chamber passing through the fuel pump, the fuel discharge duct comprehensive of
a calibrated orifice, the cooling channels, the Venturi duct, the turbine, the turbine
discharge duct and finally the FSOV.
The orifice diameter present in the discharge duct is determined during the
engine calibration depending upon the discharge coefficient value of the individual
components. The calibration orifice is represented by a pressure drop equation
with sonic speed limitation implemented in the Junction component. The value
of the loss coefficient is calculated to get the desired value for the pressure drop
according to Section 3.2.2.
The Venturi pipe downstream of the cooling jacket is intended primarily to
help provide stable thrust control using a turbine bypass valve rather than an
in-line valve. The RL-10 Venturi is apparently choked during engine start but not
at the normal operating conditions. The model presented by Binder [15] made
use of a performance map based on the inlet-to-exit pressure ratio. This model
was quite simple and needed the implementation of a inertial damping logic to
perform shut-down simulations. In this work a complete profile of the Venturi
duct has been reconstructed instead, based on the data provided by Binder and
direct measurements of the component, in order to be compliant with inlet, the
throat and the diffuser exit diameters. The duct profile has been implemented in a
pipe component which has been accordingly discretised to have a fine mesh in
proximity of the throat.

158
6. Integrated Validation: RL-10 design and analysis

Name Value Units


Inlet Diameter 0.054 [m]
Throat Diameter 0.029 [m]
Outlet Diameter 0.054 [m]
Tube Length 0.75 [m]
Number of nodes 17 [-]
∆Pref 1.2 [bar]
ṁref 2.78 [kg/s]

Figure 6.9.: Venturi nozzle profile Table 6.2.: Venturi geometrical data

For the all valves installed in the fuel line a loss coefficient is used for the
pressure loss calculation. The determination of the loss coefficient is described in
Equation 6.15:
2 · ∆Pref · ρref · A2
ζ= (6.15)
ṁ2ref

The volumes associated with this device are implemented as extra volumes in
the components upstream and downstream the valve. Due to the small dimensions
of the valve, the code of the component does not need the implementation of
a heat transfer model, which is neglected. The input parameters are shown in
Table 6.3.

Oxidiser line set


The oxidiser line setup enables the oxygen flow from the OINV to the combus-
tion chamber passing through the oxidiser pump, the OCV valve and the oxidiser
discharge duct.
Particular attention has been paid to the OCV since this component represent the
most complex valve to be modelled. As already mentioned the OCV is composed
of two orifices and its main orifice is actuated by differential pressure between
the oxidiser pump. The valve model presents two valves in parallel and they
have been calibrated in order to achieve the OCV performances during engine
start, steady state and shut-down phases. The pressure loss coefficients have been
evaluated using the same equation as for the other valves (Eq. 6.15). The input
parameters are shown in Table 6.4.

159
6. Integrated Validation: RL-10 design and analysis

Name Value Units


FINV
Reference Flow Area 0.0041 [m2 ]
ζ 1.0029 [-]
∆Pref 0.0325 [bar]
ṁref 2.8 [kg/s]
τres 17 [ms]
FCV-1
Reference Flow Area 1.44E-6 [m2 ]
ζ 2.7778 [-]
∆Pref 36.601 [bar]
ṁref 0.0195 [kg/s]
τres 10 [ms]
FCV-2
Reference Flow Area 8.6E-7 [m2 ]
ζ 2.7778 [-]
∆Pref 73.835 [bar]
ṁref 0.0163 [kg/s]
τres 10 [ms]
TCV
Reference Flow Area 1.66E-6 [m2 ]
ζ 1 [-]
∆Pref 16.030 [bar]
ṁref 0.072 [kg/s]
τres 10 [ms]
FSOV
Reference Flow Area 0.0021 [m2 ]
ζ 1.397 [-]
∆Pref 2.634 [bar]
ṁref 2.78 [kg/s]
τres 10 [ms]

Table 6.3.: Fuel line valves parameters

160
6. Integrated Validation: RL-10 design and analysis

Name Value Units


OINV
Reference Flow Area 0.0041 [m2 ]
ζ 1.054 [-]
∆Pref 0.055 [bar]
ṁref 14.207 [kg/s]
τres 17 [ms]
OCV-1
Reference Flow Area 3.44E-4 [m2 ]
ζ 0.88 [-]
∆Pref 5.84 [bar]
ṁref 13.006 [kg/s]
τres 10 [ms]
OCV-2
Reference Flow Area 5.5E-5 [m2 ]
ζ 2.685 [-]
∆Pref 5.84 [bar]
ṁref 1.201 [kg/s]

Table 6.4.: Oxidiser line valves parameters

161
6. Integrated Validation: RL-10 design and analysis

6.3. Subsystem simulation: validation at nominal conditions

Each component of the RL-10A-3-3A engine has been previously simulated as a


stand-alone component to validate its behaviour at steady state conditions, then
they have been grouped in several subsystems:

- Turbopump assembly

- Thrust chamber and cooling jacket

- Oxidiser pipe line

- Fuel pump to cooling jacket pipe line

- Cooling jacket to turbine pipe line

- Turbine to chamber pipe line

All subsystem models have then been connected together to create the complete
RL-10A-3-3A engine model (see Figure 6.10. Two different configurations of the
engine model parameters have been adopted: the first one to match the engine
nominal operation point and a second one to match the ground test results. What
differs from the two configuration is the temperature and pressure at the inlet
of the pumps and the trimming of the OCV valve to obtain the desired Mixture
Ratio.
Nominal operation point has been considered for the steady-state performance
prediction. Flight data have not been considered in this comparison because insuf-
ficient data exist to determine the mixture ratio and trim position of the oxidiser
control valve (OCV). Table 6.5 shows relative performance predictions of the tran-
sient model at steady state conditions. Where available, experimental values at
the end of the transient phase have been used as reference [15]; other performance
parameters have been compared at their nominal operating condition [15, 4, 58].

162
6. Integrated Validation: RL-10 design and analysis

Name Description Value Error


Pcc [bar] Chamber pressure 32.696 -0.16%
M R [-] Mixture Ratio 5.025 -0.58%
mcc,ox [kg/s] LOX chamber mass flow 14.102 -0.38%
mcc,f u [kg/s] H2 chamber mass flow 2.806 +0.77%
mcc,t [kg/s] Total chamber mass flow 16.908 +0.09%
∆Pcj [bar] Cooling jacket pressure drop 13.877 +0.136%
Tin,t [K] Turbine inlet temperature 204.235 -4.31%
Πtt [-] Turbine pressure ratio 1.403 -0.33%
ωt [rpm] Turbine rotational speed 31541 -0.015%
τt [N·m] Turbine torque 180.47 -1.3%
ṁt [kg/s] Turbine mass flow 2.784 -1.35%
Wt [kW] Turbine power 596.103 -1.31%
ṁp,ox [kg/s] LOX pump mass flow 14.102 -0.38%
ωp,ox [rpm] LOX pump rotational speed 12949 -0.015%
τp,ox [N·m] LOX pump torque 63.476 -5.81%
Wp,ox [kW] LOX pump power 86.082 -5.8%
ṁp1,f u [kg/s] H2 1st stage mass flow 2.842 -0.77%
ṁp2,f u [kg/s] H2 2nd stage mass flow 2.822 -0.77%
ωp,f u [rpm] H2 rotational speed 31541 -0.015%
τp1,f u [N·m] H2 1st stage torque 73.158 +2.17%
τp2,f u [N·m] H2 2nd stage torque 79.626 +0.16%
Wp1,f u [W] H2 1st stage power 241.645 -0.32%
Wp2,f u [W] H2 2nd stage power 263.01 -0.77%
T hrust [kN] Engine thrust 72.352 -1.42%
Isp [s] Engine specific impulse 440.751 -0.69%

Table 6.5.: RL-10A-3-3A engine system output data

163
6. Integrated Validation: RL-10 design and analysis

Figure 6.10.: RL-10A-3-3A schematic model

164
6. Integrated Validation: RL-10 design and analysis

6.4. RL-10 Engine start-up

6.4.1. Description of the start-up sequences

The RL-10 engine starts by using the pressure difference between the fuel tank
and the nozzle exit (upper atmospheric pressure), and the ambient heat stored in
the metal of the cooling jacket walls. The engine “bootstraps” to full-thrust within
two seconds after ignition.
Before engine start, FINV and OINV valves are opened and propellants are
allowed through the fuel pump for five seconds (cooled to prevent cavitation at
engine start) and through the LOX system for nine seconds. This “pre-start” flow
consumes approximately 10 kg of oxygen and 2.7 kg hydrogen [115]. The fuel
FCV-1 and FCV-2 valves (see Figure 6.2) are open and the main shut-off valve
(FSOV) is closed. The fuel flow is vented overboard through the cool-down valves
and does not flow through the rest of the system; the latent heat in the metal of
the combustion chamber cooling jacket is therefore available to help drive the
start transient. The oxidiser pump is pre-chilled by a flow of oxygen, which passes
through the Oxidiser Control Valve (OCV) and is vented through the combustion
chamber and nozzle.
A typical plot of the valve positions during engine start is shown in Figure 6.11.
To initiate start, the FSOV is opened and the fuel-pump discharge cool-down
valve (FCV-2) is closed. The interstage cool-down valve (FCV-1) remains partially
open in order to avoid stalling of the fuel pump during engine acceleration. The
pressure drop between the fuel inlet and the combustion chamber drives fuel
through the cooling jacket, picking up heat from the warm metal. This pressure
difference also drives the heated fluid through the turbine, starting rotation of
the pumps, which drive more propellant into the system. At start, the OCV also
closes partially, restricting the flow of oxygen into the combustion chamber. This
is done to limit chamber pressure and ensure a forward pressure difference across
the fuel turbine after ignition of the thrust chamber.
Ignition of the main combustion chamber usually occurs approximately 0.3
seconds after the main-engine start signal (t = 0) is given (for first-burns). The
ignition source is an electric spark, powering a torch igniter. The ignited com-
bustion chamber provides more thermal energy to drive the turbine. As the

165
6. Integrated Validation: RL-10 design and analysis

Figure 6.11.: RL-10A-3-3A Valve schedule for Start-up Simulation [15]

166
6. Integrated Validation: RL-10 design and analysis

turbopumps accelerate, engine pneumatic pressure is used to close the interstage


cool-down valve completely and open the OCV at pre-set fuel and LOX pump
discharge pressures. The OCV typically opens very quickly and the resultant
flood of oxygen into the combustion chamber causes a sharp increase in system
pressures. During this period of fast pressure rise, the thrust control valve (TCV)
is opened, regulated by a pneumatic lead-lag circuit to control thrust overshoot.
The engine then settles to its normal steady-state operating point.

6.4.2. Start transient

The results of start transient simulations (“Simulation” on the plots) were com-
pared with measured data of a single ground test first-burn (P2093 Run 3.02 -
Test 463, “Ground Test [3]” on the plots) [15] and with the simulation results of
a previous work (“xx_sym [3]” on the plots) performed by a NASA team [15].
Since no detailed initial conditions along the engine were available, a simulation
of the pre-start phase was necessary to obtain reasonable initial conditions for the
engine start.
The inlet pressure and temperature used in the model are coming from nominal
operation conditions. Another important variable is the cooling jacket initial
temperature, that has been set at 300 K before the pre-start phase occurs. After the
pre-start simulation the cooling jacket wall temperature decreased around 240 K.
It is clear that the cooling jacket wall temperatures have a great importance since
they help to determine the engine start capability. The cooling jacket manifold has
a lower temperature than the cooling jacket because it is partially filled by gaseous
hydrogen that has not vented overboard via the fuel discharge valve (FCV-2).
In the simulation the ignition occurs when the propellant mixture ratio inside
the combustion chamber reaches a value lower than 30 (as shown in Ref. [114]),
at around 0.3 seconds as in the ground test.

Figure 6.13 (a) shows the comparison between measured and predicted cham-
ber pressure. The model matches the measured time-to-accelerate to within
approximately 92 milliseconds (the “time-to-accelerate” is defined here as the time
from 0 seconds at which the chamber pressure reaches 13.79 bar (200 psia)). The
very first pressure rise at ≈0.3s represents the chamber ignition as mentioned

167
6. Integrated Validation: RL-10 design and analysis

above. The chamber pressure shows a “plateau” until the OCV opens. After the
OCV opening, the chamber pressure rises very quickly and then stabilizes to the
steady state condition thanks to the TCV valve closed loop control.
The presence of small oscillations evident in the test data are due to oscillations
of the TCV servo-mechanism. Such a mechanism is absent in the model so no
oscillations occur. To obtain a reasonable chamber pressure profile the TCV
opening sequence has been modified (see Figure 6.12) using the opening sequence
obtained from a dynamic model of the TCV valve as a guideline [113]. This new
opening sequence uses as a period the time that the dynamical model takes to
reach the steady condition after a sequence of oscillations. With this new opening
schedule the difference with the previous model of the RL-10A-3-3A transient
start-up [15] is remarkable.

Figure 6.12.: Valves opening sequence adopted in the simulation

168
6. Integrated Validation: RL-10 design and analysis

In Figure 6.13 (b) the LOX pump rotational speed is shown: the simulation
result is in good agreement with experimental result. The difference in the rate of
change of the pump speed between the simulation and experiment may be due to
the uncertainty in the pump inertia distribution.

Figures 6.14 (a, b) depict the LOX pump inlet and outlet pressures evolution.
The simulation exhibits some sharp transient before reaching steady-state condi-
tions; these seem to be due to fluid compressions and phase changes that occur
when the OCV suddenly opens. These transients are steeper than the measured
data probably because the dynamic behaviour of the OCV valve plays an impor-
tant role in the fluid dynamics during pressure rise. The OCV valve component
indeed, has been modelled by an open-loop control logic, while the real component
has pressure controlled mechanism function of the inlet and outlet pump pressures.

For the same reason an oxygen mass flow peak is present in the simulation
at the inlet of the engine, as it is illustrated in Figure 6.15 (a).
Figure 6.15 (b) shows the fuel inlet mass flow trend; as for the chamber pressure,
also for the measured hydrogen mass flow the evident oscillations are explained
by the oscillations of the TCV close-loop control mechanism. Unfortunately on
the fuel side no turbopump measured data are available so no comparison has
been possible between the simulated and experimental results.

The last two measured points in the engine were the pressure at the Venturi
inlet and the temperature at the Turbine inlet. The first one is illustrated in Fig-
ure 6.16 (a): it is evident that the simulation evolution is in a very good agreement
with experimental data, but also here can not represent the pressure oscillations
due to the TCV valve.
The turbine inlet temperature trend is depicted in Figure 6.16 (b): the tempera-
ture value at time = 0s represents the initial condition obtained after the simulation
of the chill-down phase, explaining the difference from the “Simulation” line and
the other two temperature plots. The temperature profile results very similar
between the two simulations due to the mixture ratio trend inside the chamber:
the engine keeps for most of the time a high mixture ratio condition but from time

169
6. Integrated Validation: RL-10 design and analysis

= 1.5 s to 1.9 s (prior to the OCV complete opening) an increase of the hydrogen
flow is noticed affecting the combustion chamber temperature.

170
6. Integrated Validation: RL-10 design and analysis

(a) Chamber Pressure

(b) LOX Pump Shaft Speed

Figure 6.13.: Transient results - part 1

171
6. Integrated Validation: RL-10 design and analysis

(a) LOX Pump Discharge Pressure

(b) LOX Pump Inlet Pressure

Figure 6.14.: Transient results - part 2

172
6. Integrated Validation: RL-10 design and analysis

(a) LOX Engine inlet mass flow

(b) Fuel Engine inlet mass flow

Figure 6.15.: Transient results - part 3

173
6. Integrated Validation: RL-10 design and analysis

(a) Venturi inlet Pressure

(b) Turbine Inlet Temperature

Figure 6.16.: Transient results - part 4

174
6. Integrated Validation: RL-10 design and analysis

6.5. RL-10 engine shut-down

6.5.1. Description of the shut-down sequence

The RL-10 engine switches off at the end of its mission, after the steady state
phase. The Fuel Shut-off Valve (FSOV) and the Fuel Inlet Valve (FINV) close as the
FCV-1 and FCV-2 valves open, allowing fuel to drain out of the system through
the overboard vents. The combustion process is soon starved of fuel and the flame
extinguishes. The Oxidiser Control Valve (OCV) and the Oxidiser Inlet valve
(OINV) begin to close next, cutting off the flow of oxygen through the engine.
The turbopump decelerates due to friction losses and drag torque created by the
pumps as they evacuate the remaining propellants from the system. A typical plot
of the valve movement during engine shut-down is shown in Figure 6.17.
During the engine shut-down, a different combination of off-design conditions
appears to exist, including pump cavitation and reverse flow. Proper simulation of
these effects is complicated by their interaction with each other. From available
test data and simulation output, it appears that as the fuel inlet valve closes and
the cool-down valves open, the pump first cavitates due to a combination of
changes in pump loading and cut-off of the inlet flow. The cavitation causes the
pump performance to degrade rapidly until the pump cannot prevent the reverse
flow of fluid as it comes backward through the cooling jacket. When the reversed
flow reaches the closed fuel inlet valve, however, extreme transients of pressure
and flow are created. Similar effects are encountered in the LOX pump during
shut-down as well.
The pump head and torque performance characteristics during, this period of
operation are, of course, not extensively documented in test data. The generic
pump characteristics found in References [130] and [25] have been used again to
extend the performance maps for cavitation and reverse flow.
The pump map extensions for engine shut-down are included in Figures 6.3,
page 149. Although the engine start-up and shut-down models use the same pump
performance maps (which should be able to cover all the pump regimes), the
cavitation and reverse flow effects also require additional modelling effort, that
has not been implemented into this model yet.

175
6. Integrated Validation: RL-10 design and analysis

Figure 6.17.: RL-10A-3-3A Valve schedule for Shut-down Simulation [15]

176
6. Integrated Validation: RL-10 design and analysis

6.5.2. Shut-down transient

The results of start transient simulations (“Simulation” on the plots) were com-
pared with measured data of a single ground test (P2093 Run 8.01 - Test 468,
“Ground Test [3]” on the plots) [15] and with the simulation results of a previous
work (“xx_sym [3]” on the plots) performed by the NASA team [15]. Differently
form the start-up simulation, the uncertainty related to the valves closing sched-
ule made necessary to slightly trim the valve sequence (few milliseconds). The
original schedule and the valves positions profile has been used as guideline. The
modified shut-down sequence is illustrated in Figure 6.18.

Figure 6.18.: Valves closing sequence adopted in the simulation

Figure 6.19 (a) illustrates the combustion chamber pressure trend. Once the
FSOV starts to close the chamber pressure decreases and this happens in both the
simulations and the experimental data, showing a good agreement between them.

177
6. Integrated Validation: RL-10 design and analysis

Figure 6.19 (b) shows predicted and measured pump speed for the Oxidiser pro-
pellant side. The discrepancies from the two models and the ground test measured
data are imputable uncertainties to exact inlet conditions and initial operating
point as well as to a precise distribution of the turbopump assemblies inertia.

Figures 6.20 (a, b) depict the LOX pump inlet and outlet pressures evolution.
Regarding the pressure at the outlet of the pump, no special features are evident.
Once the FSOV valve starts to close, the outlet pump pressure decreases because
of the minor power delivered by the turbine.
Figure 6.20 (b) illustrates the inlet pressure instead. From the measured data we
see an initial pressure decrease due to the pump conditions and then a recovery in
the pressure to the complete closure of the OINV valve. This behaviour is barely
reproduced by the simulation because of the lack of a cavitation model in the
pump, hence the final pressure decrease is not as evident as in the experiment.

The engine propellant mass flows are depicted in Figures 6.21 (a, b), for the
oxidiser and the fuel respectively. The oxygen mass flow behaviour (Figure 6.21
(a)) is mainly function of the pump behaviour; it is interesting to underline that
from analyses performed varying the opening/closing time of the valves, the role
of the FCV valves becomes much more evident. The opening of the FCV valves
decreases the turbine power, thus decreasing the propellant mass flow rate in the
system in order to avoid mass flow rate surges of oxygen at the FSOV closure.
In the end, the complete shut-off of the OINV valve extinguishes the propellant
flow rate.
A more complex profile is present in the fuel flow plot as shown in Figure 6.21
(b): at the beginning of the shut-down phase the hydrogen mass flow at the engine
inlet increases because of the opening of the FCV valves. Then the closure of the
FSOV and of the FINV valve determine the mass flow shut-off. The simulation
reproduces correctly what happens at inlet of the engine, even though the amount
of mass flow venting through the FCV valves results too high determining a
higher peak at the inlet respect to the one observed in the ground test. Another
interesting point to be mentioned is that differently from the NASA results the
“Simulation” line does not show any reverse flow at the inlet as well as the experi-

178
6. Integrated Validation: RL-10 design and analysis

mental data.

The RL10 shut-down model has captured many interesting effects that occur
during shut-down. In Figure 6.22 (a), for example, the measured data show a
characteristic dip, rise and then falloff in the fuel venturi upstream pressure. This
features is caused by the dynamic interaction of the fuel pump cool-down valve
opening and main fuel shut-off valve closing. It is very likely that the absence
of this peculiar behaviour inside our model is due to a not perfectly precise
synchronization of the fuel valves closing schedules.
In Figure 6.22 (b), the jump in pump inlet pressure is due, in part, to reverse
flow through the fuel pump. As already mentioned, a cavitation model for pump
performance deterioration is not implemented so the pressure peak does not rise
in the simulation result.
From inspection of the plots, it appears that there are still unresolved differences
between the predicted and measured engine deceleration rates. The discrepancies
can be tracked down due two main causes: first, the time scales of the shut-down
processes are much smaller than the one from the start-up transient, and second
the complex phenomena such as cavitation and blade to fluid interaction that are
not taken into account into the present model.

179
6. Integrated Validation: RL-10 design and analysis

(a) Chamber Pressure

(b) LOX Pump Shaft Speed

Figure 6.19.: Shut-down results - part 1


180
6. Integrated Validation: RL-10 design and analysis

(a) LOX Pump Discharge Pressure

(b) LOX Pump Inlet Pressure

Figure 6.20.: Shut-down results - part 2


181
6. Integrated Validation: RL-10 design and analysis

(a) LOX Engine inlet mass flow

(b) Fuel Engine inlet mass flow

Figure 6.21.: Shut-down results - part 3


182
6. Integrated Validation: RL-10 design and analysis

(a) Venturi inlet Pressure

(b) Fuel Pump Inlet Pressure

Figure 6.22.: Shut-down results - part 4


183
6. Integrated Validation: RL-10 design and analysis

6.6. Dynamic Response Analysis

In this section the dynamic response of the entire engine system to valve per-
turbations is investigated and illustrated. The thrust control valve TCV and the
oxidiser control valve OCV have been throttled in the range of ±10% around their
nominal aperture ratio by use of a step function.
The main objective is to understand the response of the engine when the valve
in charge of the thrust control (TCV) or the valve in charge of mixture ratio
definition are throttled.
The simulations performed have as initial operating condition the nominal
steady state point.

Thrust Control Valve throttling :


The thrust control Valve (TCV) has been operated in order to give an instantaneous
aperture ratio signal from 0.3 to 0.27 and from 0.3 to 0.33, that is a signal equal to
±10% of its nominal value.
When the TCV valve is operated, the main objective is to modify the thrust of
the engine. Closing or opening the bypass TCV valve we increase or decrease
respectively the fuel mass flow rate into the turbine, varying the delivered power.
Figures 6.23 to Figures 6.26 show how the main engine subsystems react to
TCV operation, while Table 6.6 summarises the main engine parameters values at
nominal operation point, the percentage difference when compared to nominal
conditions, the value of τR the response time required to reach 90% of the final
value at steady state conditions and some comments about the characteristics of
the curves shown hereafter (“Sym” stands for a symmetric trend of the variable
respect to ±10% of the valve opening, “Os” stands for Overshoot and “Rev” stands
for Reverse, that is when a variable presents a change in its trend).
Since the bypass valve has a very little cross section area, a 10% modification on
its aperture ratio does not affect so much the engine parameters. It is interesting
to underline that, since the TCV operation has a direct influence on the turbine
that control both propellants’ lines, the percentage variation with +10% or -10% is
almost the same for each parameter in both lines.

184
6. Integrated Validation: RL-10 design and analysis

As expected, reducing the aperture ratio of the TCV valve we find that chamber
pressure increases, mixture ratio shows a very slight increase as well as the
discharge pumps pressures (see Figures 6.24 (a,b)) and the injected propellants
into the chamber (see Figures 6.25 (a,b)). Please note that the fuel mass flow rate
injected into the combustion chamber decreases at the very beginning instants
after the valve aperture ratio reduction; then the system starts to react to the
increase of turbine power and subsequently we see the hydrogen flow increasing
into the chamber. The time required to the system to react is defined as the τR
parameter. From Table 6.6 it is clear that τr is between 0.3 and 0.4 s, one order
of magnitude the response time of the valve. This time is mainly function of the
inertia of the turbopump assembly and of the length of the pipes.
Another interesting point to underline is that for a modification of ±10% of the
valve aperture ratio corresponds a percentage variation in the valve mass flow
rate of almost the same quantity (see the line in Table 6.6).

185
6. Integrated Validation: RL-10 design and analysis

Variable Nominal ∆-10% [%] ∆+10% [%] τR [s] Notes


Pcc 32.87 +0.243 -0.243 0.36 Sym
Tcc 3252.19 +0.069 -0.068
MR 5 +0.189 -0.19 0.09 Os
ṁox 14 +0.305 -0.301 0.352 - 0.335 Sym
ṁf u 2.8 +0.116 -0.116 0.407 - 0.37 Rev
ṁcc,t 16.8 +0.273 -0.274
Pinj,ox 37 +0.288 -0.286
Pinj,f u 36.67 +0.224 -0.224
Wt 589.99 +0.352 -0.352
τt 179.22 +0.242 -0.242
ωt 31437 +0.11 -0.111 Sym
Πtt 1.4 +0.023 -0.023
Pout,ox 45.17 +0.345 -0.342 0.358 - 0.348 Sym
τp,ox 62.89 +0.356 -0.353
ωox 12906.9 +0.111 -0.111
Wp,ox 85 +0.467 -0.463
Pout,f u 73.58 +0.202 -0.202 0.355 - 0.384 Sym
τp1,f u 72.66 +0.218 -0.218
τp2,f u 79.12 +0.226 -0.227
Wp1,f u 239.21 +0.329 -0.329
Wp2,f u 260.47 +1.699 -1.247
∆POCV 5.61 +0.607 -0.597
ṁOCV −1 12.82 +0.304 -0.3
ṁOCV −2 1.18 +0.305 -0.3
∆PT CV 16.08 +0.293 -0.293
ṁT CV 0.02 -9.798 +9.752

Table 6.6.: Engine dynamic response to TCV ±10% operation

186
6. Integrated Validation: RL-10 design and analysis

(a) Chamber Pressure

(b) Chamber Mixture Ratio

Figure 6.23.: TCV throttle results - part 1

187
6. Integrated Validation: RL-10 design and analysis

(a) LOX Pump Discharge Pressure

(b) Fuel Pump Discharge Pressure

Figure 6.24.: TCV throttle results - part 2

188
6. Integrated Validation: RL-10 design and analysis

(a) LOX chamber inlet mass flow

(b) Fuel chamber inlet mass flow

Figure 6.25.: TCV throttle results - part 3

189
6. Integrated Validation: RL-10 design and analysis

(a) Turbine Mass Flow

(b) Turbine Shaft Speed

Figure 6.26.: TCV throttle results - part 4

190
6. Integrated Validation: RL-10 design and analysis

Oxidiser Control Valve throttling :


The Oxidiser Control Valve (OCV) has been operated in order to give an instanta-
neous aperture ratio signal from 0.745 to 0.6705 and from 0.745 to 0.8195, that is a
signal equal to ±10% of its nominal value.
When the OCV valve is operated, the main objective is to modify the mixture
ratio of the engine. Closing or opening the control valve OCV we increase or
decrease respectively the oxidiser mass flow rate into the oxidiser line of the
engine, varying the delivered propellant amount into the chamber.
Figures 6.27 to Figures 6.30 show how the main engine subsystems react to OCV
operation, while Table 6.7 summarises the main engine parameters values at nom-
inal operation point, the value of τR and some comments about the characteristics
of the curves shown hereafter.
Since the OCV valve operates in the oxidiser line, the behaviour of the entire
system results more complex and a deeper investigation of how system reacts is
needed. The fuel side of the engine system is involved by the valve operation
indirectly. A decreasing of the OCV aperture ratio generates a increase of the
valve resistance leading to a decrease of the oxidiser mass flow rate as well as
an increase of the LOX pump head rise. Both contrasting phenomena lead to a
decrease of the shaft pump required torque on the oxidiser side determining an
acceleration of turbopump subsystem and involving in this way the fuel side of
the engine, in which the hydrogen mass flow shows an increment. For these many
reasons the combustion chamber pressure shows a slight percentage increment.
The opposite behaviour is shown when we open of 10% the OCV valve. As
expected, the resistance of the valve is lower, the shaft pump torque goes up, hence
the oxygen mass flow rate increases while the hydrogen flow shows a decrease
because of the minor power delivered by the turbine to the pumps.

191
6. Integrated Validation: RL-10 design and analysis

Variable Nominal ∆-10% ∆+10% τR [s] Notes


Pcc 32.87 +0.544 +0.15 0.51 Rev
Tcc 3252.19 -1.226 +2.046
MR 5 -3.533 +6.801 0.474 - 0.429 Os
ṁox 14 -0.626 +2.385 0.53 - 0.59 Os
ṁf u 2.8 +3.013 -4.135 0.497 - 0.47 Os
ṁcc,t 16.8 -0.021 +1.298
Pinj,ox 37 +0.331 +0.71
Pinj,f u 36.67 +0.799 -0.351
Wt 589.99 +4.606 -4.691
τt 179.22 +3.091 -3.707
ωt 31437 +1.47 -1.022 Asym
Πtt 1.4 +1.048 -1.058
Pout,ox 45.17 +2.647 -0.647 0.420 - 0.449 Asym-Rev
τp,ox 62.89 +1.503 -0.217
ωox 12906.9 +1.471 -1.022
Wp,ox 85 +2.995 -1.237
Pout,f u 73.58 +2.12 -2.266 0.435 - 0.413 Asym
τp1,f u 72.66 +2.174 -3.694
τp2,f u 79.12 +4.45 -4.858
Wp1,f u 239.21 +3.676 -4.678
Wp2,f u 260.47 +5.986 -5.83
∆POCV 5.61 +19.672 -12.06
ṁOCV −1 12.82 -1.553 +3.178
ṁOCV −2 1.18 +9.386 -6.202

Table 6.7.: Engine dynamic response to OCV ±10% operation

192
6. Integrated Validation: RL-10 design and analysis

(a) Chamber Pressure

(b) Chamber Mixture Ratio

Figure 6.27.: OCV throttle results - part 1

193
6. Integrated Validation: RL-10 design and analysis

(a) LOX Pump Discharge Pressure

(b) Fuel Pump Discharge Pressure

Figure 6.28.: OCV throttle results - part 2

194
6. Integrated Validation: RL-10 design and analysis

(a) LOX chamber inlet mass flow

(b) Fuel chamber inlet mass flow

Figure 6.29.: OCV throttle results - part 3

195
6. Integrated Validation: RL-10 design and analysis

(a) Turbine Mass Flow

(b) Turbine Shaft Speed

Figure 6.30.: OCV throttle results - part 4

196
7. Conclusions

Transient phenomena in liquid rocket engines, ranging from combustion high


frequency instabilities to water hammer effects in the feed lines, and which poten-
tially result in system failures, drive the necessity to dedicate special attention to
transient phases.
Concentrating on the behaviour of only one component is however not sufficient
to understand how components affect each other during such phases, what is their
impact on system frequencies, and how this interaction may lead to a failure.
The simulation of the complex flow behaviour in engine components and
components assemblies is therefore required. Models allowing the examination of
detailed component flow behaviour are based on the equations of conservation
of mass, momentum, and energy, and vary widely in their complexity and in the
computational time each requires. An intelligent simplification of the underlying
processes allows to reduce the governing partial differential equations to ordinary
differential equations, which no longer require complex solution methods thus
allowing much faster computational times.
The development of model capable of simulating in a more accurate way with
respect to previous models liquid rocket engine components and propulsion sys-
tems resulted in the work performed in this thesis. Implemented in the ESPSS
library, they can simulate the major liquid rocket engine components: pipes,
valves, injector domes, injectors, turbopumps, combustion devices and nozzles.
For the creation of a steady state library, each component has been tested to
validate its behaviour at component level and then in a more complex system. The
models developed and improved for transient analyses have been validated either
with CFD numerical test cases or experimental results. Each one of them, tested
in system, displays its own dynamics and characteristics which when integrated
in a more complex component assemblies are seen to interact.

197
7. Conclusions

A new library for steady state applications has been presented and validated.
The library enables to perform in a fast and reliable way design and parametric
analyses of liquid propulsion systems. The present work has described a complete
set of components able to perform dimensioning design studies and off-design
analyses of liquid propulsion systems. A gas generator and an expander cycle
have been chosen to validate the design capability of the steady state library.
The resulting designs have been compared with actual liquid rocket engine test
data. The steady state results when compared with nominal values show a good
agreement as a proof of the accuracy of the library.

The injector head model has been tested with a realistic test case. The new
structure of the injector dome allows to take into account the strong interaction
between the combustion chamber, the propellants in the injector dome and the
injector dome walls, evaluating the transient heat fluxes which rise during the
ignition of an engine. The implementation of this new model has a fundamental
importance for the correct representation of the two-phase flow inside the injector
dome and the mass flow evolution during start-up and shut-down.

Hot-gas-side heat fluxes in combustion devices are now described in a more


detailed way, making use of different correlations for the evaluation of the hot-
gas-side heat transfer coefficient hc . The presence of different correlations and the
possibility to choose different and fine tuned correction factors allows the study
of a propulsion subsystem varying the heat flux behaviour calculation.

Representation of thermal stratification inside high aspect ratio cooling chan-


nels, and its development along transient conditions required a modification of the
basic one-dimensional equations for pipes, combining semi-empirical correlations
and a Quasi-2D approach in order to save computational time and therefore keep
the model useful for system simulation purposes.
The quality and the robustness of this model has been proved first comparing
its results with CFD numerical test case, and then with experimental results from
a test campaign especially performed for the evaluation of thermal stratification

198
7. Conclusions

in this kind of cooling channels.

The modelling capabilities at system level have been deeply demonstrated with
the development and the creation of a model for the an entire liquid propellant
rocket engine, the RL-10A-3-3A. The construction of a model to simulate start-up
and shut-down phases of this engine required the investigation of all the main
critical aspects which occur during transient phases for all the components that
assemble the engine. Simulations for the engine pipelines, throttle and regulation
valves, turbine and pump assemblies, cooling channels and combustion chamber
have been performed to verify the correct behaviour of the components and of the
subsystems when compared with actual data.
Comparison of the transient behaviour of the engine during ground test and
model predictions is very satisfactory. Although many uncertainties affect the
transient simulation (such as valve discharge coefficient uncertainties, running
shaft torque, oxidiser control valve behaviour, initial conditions uncertainties etc.)
the model correctly reproduces the main phenomena occurring during transients,
such as combustion, heat transfer, turbopump operation phase change, valve
manoeuvering and pressure drops, as well as the thermodynamic behaviour of
the fluids. Two phase flow effects in the engine are also well estimated. Moreover
the RL-10A-3-3A model accurately predicts the engine time-to-accelerate when
compared to ground test data.

The models developed and the simulations performed at component level and at
system level, and the understanding gathered during the analysis of the transient
phases of the RL-10 engine stimulate for further improvements and developments
to increase the reliability of such a tool for prediction and evaluation of the
transient phases of a liquid rocket engine propulsion system.
Developments could include the following:

• Taking into account the injected liquid phase into the combustion chamber

• Implementation of a chemical kinetics algorithm for finite rate combustion


model

199
7. Conclusions

• Development of a film cooling model into the combustion chamber

• Implementation of a fully transient model for the nozzle component

• Inclusion of heat transfer and mass capacitance effects into turbopump


models

• Inclusion of a cavitation model for pumps

In conclusion, the deep investigation in the characteristic problems that may


occur during transient phases of a liquid rocket engine and the work performed
in this thesis have brought to the development of more accurate and complex
models to evaluate peculiar phenomena inside liquid propulsion systems and the
identification of additional work in order to have in the future a very reliable tool
for the prediction of liquid rocket engines start-up and shut-down phases.

200
Appendices

a
A. Implementation of Up-wind Roe Scheme

A.1. Governing equations

Here is recalled the set of governing equations that will be used in the fluid_flow_1d
library of EcosimPro. They are derived from the 1D Navier-Stokes equations,
using the conservative set of variables u = (ρ, ρu, ρE) :

∂ρ ∂p ∂ρu
+ κw ρ + = 0
∂t ∂t ∂x
∂ρu ∂(ρu2 + p)
+ = −Fw − ρg (A.1)
∂t ∂x
∂ρE ∂ρuH
+ = Qw
∂t ∂x

where :

• The geometry can be quasi-1D : cross-section A can smoothly vary, so


the set of variables should change to uA = (ρA, ρuA, ρEA). Using these
variables an extra term accounting for the cross-section variation arises
in the momentum equation. This particular case of variable cross-section
is studied in the section (A.3.3). The general case is considered to be the
constant cross-section formulation (A.1);

• In order to simulate accurately the water hammer effect, the wall compress-
ibility κw must be taken into account, through an extra term1 in the mass
conservation equation. The derivation of this term can be found in the
fluid_flow_1d Manual [43];

• The gradient of shear stresses is represented in the momentum equation as


a source term Fw including all possible pressure losses in the component;
1
implemented as a source term in a first approximation.

b
A. Implementation of Up-wind Roe Scheme

• The work of shear stresses and external forces is neglected;

• There is one heat source representing the heat transfer with the wall Qw .

The system (A.1) of mass, momentum and energy conservation equations can be
applied to either :

• a one component, one phase fluid : some gas or liquid;

• a one component, two-phase fluid : the fluid can undergo some phase
change. In this case all the variables, as well as thermodynamic and transport
coefficients involved in the system (A.1), correspond to either the gas, the
liquid, or the 2-phase mixture, depending on the operating conditions. More
details are found in the fluid_flow_1d Manual of EcosimPro.

A.1.1. 4-equation subset

The system of 3 equations above must also be extended to the case of a mixture of
two components, for which case the first one can be either one phase or two-phase,
and the second one is always a non-condensable gas. Here an asymmetric formu-
lation is chosen, thus using the conservative set uasym = (ρ, ρnc , ρu, ρuE) rather
than usym = (ρ1 , ρ2 , ρu, ρuE). The resulting two-component set of governing
equations is :

∂ρ ∂P ∂ρu
+ κw ρ + = 0
∂t ∂t ∂x
∂ρnc ∂ρnc u
+ = 0
∂t ∂x
∂ρu ∂(ρu2 + p)
+ = −Fw − ρg (A.2)
∂t ∂x
∂ρE ∂ρuH
+ = Qw
∂t ∂x

Here these equations govern the conservation of the mixture mass, non-condensable
mass, mixture momentum and mixture energy, respectively. Details about the mix-
ing rules can be found in the fluid_flow_1d Manual [43] and are also recalled
in the next chapters.

c
A. Implementation of Up-wind Roe Scheme

A.2. Numerical concepts

Roe’s flux difference splitting (FDS) method with the MUSCL-TVD scheme is
the most famous numerical scheme applied to enhance the numerical stability,
especially for steep gradients in density and pressure near the gas-liquid interface.

We intend to use that compressible scheme for all models under consideration,
because even the liquid flows are influenced here by compressibility effects. Some
significant test cases will be performed in order to verify the accuracy of this
method.

In this section the Roe scheme is described, then the MUSCL reconstruction
methodology is given. Afterwards, a few words on the variable cross-section
formulation and consequences are given.

A.2.1. Roe’s numerical scheme

Considering hereafter the set of n general equations in matrix conservative form,


describing the behaviour of a fluid:

∂ u ∂ f(u)
+ = S(u) (A.3)
∂t ∂x

Equivalently, in quasi-linear form

∂u ∂u
+ J(u) · = S(u) (A.4)
∂t ∂x

where u is a set of n conservative variables, f(u) is the conservative flux, S(u) is


the source vector containing all terms that cannot be expressed in conservative
form, and J(u) is the Jacobian of the system defined by :

∂ f(u)
J(u) ≡ (A.5)
∂u

We focus our interest here to the convective part of (A.3), namely

∂ u ∂ f(u)
+ =0 (A.6)
∂t ∂x

d
A. Implementation of Up-wind Roe Scheme

or equivalently
∂u ∂u
+ J(u) · =0 (A.7)
∂t ∂x
which form a well-defined initial-valued hyperbolic problem provided that the
Jacobian matrix J(u) has real eigenvalues and that some initial value u(x, 0) =
u0 (x) is given.

A.2.2. Approximate Riemann Solver

Several upwind differencing schemes, based on a cell-centered (collocated) 1D


finite volume formulation, have been developed to solve the conservation equa-
tions (A.6) in each control cell i = [i − 12 , i + 12 ] in integral form. Therefore, the
spatial evolution of the conservative variables u is piece-wise constant (constant
on each cell i), and the flux expressions have to be evaluated at each cell interface
i + 21 = [i, i + 1] = [L, R] (and similarly at i − 12 ), as the discretisation of the flux
is :
∂ f fi+ 1 − fi− 1
2 2
≈ (A.8)
∂x i
∆x
The interface separates the ‘left’ state denoted by uL and the ‘right’ state denoted
by uR . This defines precisely a non-linear Riemann problem. Starting from Go-
dunov’s original scheme [54], those schemes attempt thus to build the solution of
(A.6) by solving a succession of Riemann problems on each cell interface of the
1D domain.

Recall that the Riemann problem is the initial-value problem for (A.7) with a
discontinuous initial condition across the interface :

uL x<0
u(x, 0) ≡ (A.9)
u
R x>0

Numerical efficiency justifies the use of a linearisation of that Riemann problem.


We concentrate here on the Approximate Riemann Solver introduced by Roe [123],
which exploits the fact that we can easily solve the Riemann problem for any
linear system of equations. So rather than solving the exact Riemann problem at
the interface, which is CPU-time consuming, we solve exactly the approximate

e
A. Implementation of Up-wind Roe Scheme

Riemann problem derived by replacing (A.7) by the local linearisation

∂u e uR , uL ) · ∂ u = 0
+ J( (A.10)
∂t ∂x

In that case, the interface flux fi+ 1 (or fi− 1 ) can be written as :
2 2

f(uL , uR ) = f(uL ) + Je− (uL , uR )(uR − uL ) (A.11)


= f(uR ) − Je+ (uL , uR )(uR − uL ) (A.12)
1  1  
= f(uL ) + f(uR ) − |J(uL , uR )| uR − uL
e (A.13)
2 2

where Je± (uL , uR ) are the positive and negative parts of the so-called Roe-matrix
J(
e uL , uR ), which must be constructed to satisfy the following set of conditions
christened by Roe as ‘Property U’:

i) J(
e uR , uL ) has real eigenvalues and a corresponding complete set of linearly
independent eigenvectors;

ii) J(
e uR , uL ) → J(u) as uL , uR → u;

iii) J(
e uR , uL ) must satisfy the relation :

∆f = J(
e uR , uL )∆u (A.14)

where the operator ∆(·) = (·)R − (·)L represents the jump in the quantity
(·) across the interface between left and right states.

Condition (i) ensures that the problem (A.10) is hyperbolic and solvable. Condi-
tion (ii) guarantees that the scheme gives satisfactory results for smooth flows.
Condition (iii) ensures that the scheme is conservative and that the approximate
solution is coincident with the exact one when the left and right states are con-
nected by a single jump satisfying the Rankine-Hugoniot conditions (accurate
shock resolution).

Initially, Roe derived the matrix J(


e uR , uL ) for a perfect gas as

J(
e uR , uL ) = J(q
e) (A.15)

f
A. Implementation of Up-wind Roe Scheme

that is, the exact Jacobian matrix but evaluated at the so-called Roe-average state
e , which itself is an arithmetic average between left and right states, but defined
q
on a parameter vector w :
 
wL + wR
Roe(q) ≡ q
e=q (A.16)
2

The Roe parameter w is defined such that u and f(u) are both quadratic functions
of w. The chosen notation intends to emphasize that the average state implies
only those variables that explicitly appear in the Jacobian matrix. It is easy to
check, in this case, that (A.15), obtained by satisfying property (iii), meets all of
the other requirements set by Property U.

Roe’s original result was dedicated to the Euler equations with perfect gases,
but it has been used by several authors to achieve a simpler way of determining
J(
e uR , uL ) for more complex systems and with other Equations of State (EoS). If
one assumes that (A.15) holds, it is possible to look immediately for the average
state q
e that satisfies property (iii) by direct substitution in (A.14) or in the eigen-
vector expansion of ∆f and ∆u. Surprisingly, an exact definition of a Roe-average
for non-perfect gases not only exists but is actually not unique. All the methods
cited above lead to a matrix J(
e uL , uR ) involving undefined coefficients, which are
the Roe-average pressure derivatives. More details can be found in the following
chapters.

Eigen decomposition

We recall that in the numerical flux expression (A.13), the absolute value of the
Roe-matrix J(e uR , uL ) is needed. For a given matrix, say the Jacobian J , the
absolute value of J is defined through its diagonalization as:

|J| = R · |Λ| · L (A.17)

where L and R are the left and right eigenmatrices respectively, and |Λ| contains
the absolute values of the eigenvalues λk of J on its diagonal. These eigenvalues

g
A. Implementation of Up-wind Roe Scheme

can be found by solving :


|J − λI| = 0 (A.18)

for λ. The absolute eigenvalue diagonal matrix is then :


 
|λ1 | 0 ··· 0
 
 0 |λ2 | · · · 0 
|Λ| =  . (A.19)
 
 .. ... .. 
 . 
0 ... 0 |λ |n

The right eigenvectors Rk forming the columns of the n × n right eigenmatrix


R = (R1 R2 · · · Rn ) are found by solving these n systems : J · Rk = λk Rk . Like-
wise, the left eigenvectors Lk forming the rows of the n × n left eigenmatrix
L = (L1 , L2 , · · · , Ln ) are found by solving these n systems: Lk · J = λk Lk . Con-
sistency imposes that L · R = I , where I is the n × n identity matrix.

In the same way, the absolute value of the Roe-matrix J(


e uR , uL ) is derived:

|J(
e uR , uL )| = R
e · |Λ|
e ·L
e (A.20)

Following property (A.15) of Roe’s scheme, stating that the linearised Jacobian
is the exact Jacobian but evaluated at some Roe-average state q e , the average
eigenvalues and eigenmatrices are the exact ones but evaluated at that Roe-
average:

L
e = L(q
e) (A.21)
R
e = R(q
e) (A.22)
Λ
e = Λ(q
e) (A.23)

Now if we project the conservative variable difference onto the right eigenvectors
ek
R
∆u = uR − uL = R
e·a
e (A.24)

h
A. Implementation of Up-wind Roe Scheme

from which one find the wave strengths

α1 , α
e = (e
a e2 , ..., α
en )t
e−1 · ∆u
= R
= L
e · ∆u (A.25)

We can now explicitly show the eigen decomposition in the diffusive term of
(A.13):

|Jei+ 1 | · ∆u = R
e · |Λ|
e ·L
e · ∆u
2

= R
e · |Λ|
e ·a e
n
ek
X
= α ek |R
ek |λ (A.26)
k=1

Recall that the numerical interface flux f(uL , uR ) is given by (A.13). Using the
above eigen decomposition, we have finally
n
1  1X
ek
f(uL , uR ) = f(uL ) + f(uR ) − ek |λ
α ek |R (A.27)
2 2
k=1

This shows that this Roe-matrix J(


e uR , uL ) is used as a characteristic-based con-
trolled amount of numerical diffusion.

A.3. Reconstruction method

A.3.1. Higher order accuracy

The order of accuracy of the scheme presented is however first order : the variables
ui are still constant within each cell. We can retrieve a higher order accuracy
scheme by reconstructing the variations ui (x) on each cell, through the MUSCL
approach (Monotone Upstream-centred Scheme for Conservation Laws).

A possibly piecewise quadratic local reconstruction of ui (x) within cell [i− 12 , i+ 12 ]


is:
∆x2 (2)
 
x − xi (1) 3ω 2
ui (x) = ui + δ ui + (x − x i ) − δ ui (A.28)
∆x 2∆x2 12

i
A. Implementation of Up-wind Roe Scheme

where ω ∈ [−1, 1] is a free parameter (see Table A.1) and δ (1/2) ui an estimation
of the first/second derivative of ui (x), respectively. Remark that the nodal value
ui (x = xi ) is not necessarily equal to ui :

ω
ui (x = xi ) = ui − δ (2) ui (A.29)
8

If we require these gradients δ (1/2) ui to depend only on adjacent cells, we have


simply:

1
δ (1) ui = 1
2 (ui+1 − ui−1 ) = (∆u1sti+ 21
+ ∆u1st
i− 12
) (A.30)
2
δ (2) ui = ui+1 − 2ui + ui−1 = ∆u1st
i+ 1
− ∆u1sti− 1
(A.31)
2 2

where the following notations for the jumps between constant values have been
introduced:

∆u1st
i− 1
≡ ui − ui−1 (A.32)
2

∆u1st
i+ 1
≡ ui+1 − ui (A.33)
2

Actually, the resolution of the approximated Riemann problem requires only the
values at the cell boundaries i ± 12 , extrapolated from (A.28):

1 ω
u+
i− 1
≡ ui (x = xi− 1 ) = ui − δ (1) ui + δ (2) ui (A.34)
2 2 2 4
1 (1) ω (2)
u−
i+ 1
≡ ui (x = xi+ 1 ) = ui + δ ui + δ ui (A.35)
2 2 2 4

where the ± superscripts denote the right/left side of the interface, respectively.
With the slope definitions (A.30)-(A.31), the extrapolated boundary values at the
interfaces i ± 12 become:

1 1
u−
i− 1
= ui−1 + (1 − ω)∆u1st
i− 3
+ (1 + ω)∆u1st
i− 1
(A.36)
2 4 2 4 2

1 1
u+
i− 1
= ui − (1 + ω)∆u1st
i− 21
− (1 − ω)∆u1st
i+ 21
(A.37)
2 4 4
1 1
u−
i+ 1
= ui + (1 − ω)∆u1st
i− 1 + (1 + ω)∆u1st
i+ 12
(A.38)
2 4 2 4
1 1
u+
i+ 1
= ui+1 − (1 + ω)∆u1st
i+ 1 − (1 − ω)∆u1st
i+ 23
(A.39)
2 4 2 4

j
A. Implementation of Up-wind Roe Scheme

The high-order reconstructed jump at the interface i + 12 is given by:

1
∆uhot
i+ 1
≡ u+
i+ 1
− u−
i+ 1
= (1 − ω)(−ui+2 + 3ui+1 − 3ui + ui−1 ) (A.40)
2 2 2 4
1 ∂ 3 ui
= − (1 − ω) 3 + O(∆x4 ) (A.41)
4 ∂x

Depending on the value of ω , different schemes and orders of accuracy can be


reached. The following table summarizes the different choices.

ω Reconstruction Order Nodal value, see (A.29) Jump value, see (A.40)
1 1
-1 linear one-sided 2nd 8 (ui−1 + 6ui + ui+1 ) 2 (−ui+2 + 3ui+1 − 3ui + ui−1 )
1
0 linear up/down 2nd ui 4 (−ui+2 + 3ui+1 − 3ui + ui−1 )
1 1 1
3 parabolic 3rd 24 (−ui−1 + 26ui − ui+1 ) 6 (−ui+2 + 3ui+1 − 3ui + ui−1 )
1
1 linear central 2nd 8 (−ui−1 + 10ui − ui+1 ) 0

Table A.1.: Different values of ω .

Some remarks on this table:

• We can see that for ω = −1, the interpolation is fully one-sided, as the
extrapolated boundary values are computed using two upstream cells;

• Using ω = 0, that extrapolation uses one upstream and one downstream cell.
Moreover, only for ω = 0 do we have a nodal value equal to the constant
value ui ;
1
• Only for ω = do we have a parabolic interpolation, and thus a third order
3
accuracy scheme. Indeed for ω = 13 , the reconstruction (A.28) is a correct
Taylor development up to the third order;

• Using ω = 1, the scheme looses its upwind behaviour, as the interpolation is


a simple arithmetic average between adjacent cells. The scheme corresponds
to a central scheme as there is no discontinuity at the interface : the jump
value is zero.

k
A. Implementation of Up-wind Roe Scheme

An example of piecewise linear reconstruction with ω = 0 is shown on Fig.A.1. We


see that without reconstruction, the jump at the interface i + 12 is ∆u1st
i , whereas
after a linear reconstruction, the jump at the interface is ∆u2nd
i .

Figure A.1.: Piece-wise linear reconstruction.

l
A. Implementation of Up-wind Roe Scheme

High-resolution scheme

A typical problem with a higher-order accurate discretisation is the spurious


oscillations, which appear in the vicinity of the non-smooth solutions (Godunov’s
theorem). The problem is solved if a combination of the first- and the higher-order
accurate discretisation is used.

Therefore, the higher order data reconstruction (A.28) is constrained through


a TVD version (Total Variation Diminishing) of this approach to retrieve a first-
order scheme near strong gradients, and able a higher-order scheme in smooths
parts of the flow. The slopes δ (1/2) ui must be limited, and therefore the extrapo-
lated boundary values have the following limited expression:

1 1
u−
i− 1
= ui−1 + (1 − ω)φ+
i− 3
∆u1st
i− 3
+ (1 + ω)φ−
i− 1
∆u1st
i− 1
(A.42)
2 4 2 2 4 2 2

1 1
u+
i− 1
= ui − (1 + ω)φ+
i− 12
∆u1st
i− 21
− (1 − ω)φ−i+ 12
∆u1st
i+ 21
(A.43)
2 4 4
1 1
u−
i+ 1
= ui + (1 − ω)φ+
i− 2
1st
1 ∆ui− 1 + (1 + ω)φ−
i+ 12
∆u1st
i+ 12
(A.44)
2 4 2 4
1 1
u+
i+ 1
= ui+1 − (1 + ω)φ+
i+ 2
1st
1 ∆ui+ 1 − (1 − ω)φ−
i+ 32
∆u1st
i+ 23
(A.45)
2 4 2 4

In these expressions, the limiting coefficients φ±


i± 1
and φ∓
i± 3
are defined as:
2 2

φ−
i− 1
= φ(r−
i− 1
) φ+
i− 1
= φ(r+
i− 1
) (A.46)
2 2 2 2

φ−
i+ 1
= φ(r−
i+ 1
) φ+
i+ 1
= φ(r+
i+ 1
) (A.47)
2 2 2 2

φ−
i+ 3
= φ(r−
i+ 3
) φ+
i− 3
= φ(r+
i− 3
) (A.48)
2 2 2 2

and the r function ‘measures’ the smoothness of the solution, as it is defined as a


ratio of consecutive variations:
∆u1st
i− 3 1
r−
i− 12
= ∆u1st1
2
= (A.49)
i− 2 r+
i− 23
∆u1st
i+ 1 1
r+
i− 12
= ∆u1st1
2
= (A.50)
i− 2 r−
i+ 21
∆u1st
i+ 3 1
r+
i+ 12
= ∆u1st1
2
= (A.51)
i+ 2 r−
i+ 23

m
A. Implementation of Up-wind Roe Scheme

The function φ can be any slope limiter, see [85] for details. It has been decided
that both MinMod φmm and SuperBee φsb limiters will be used and tested. They
are defined as:

φmm (r) = max [0, min(1, r)] (A.52)


φsb (r) = max [0, min(2r, 1), min r, 2] (A.53)

In that way, the scheme is second- or possibly third-order accuracy in space in


the smooth parts of the flow, and reduces to first order where strong gradients
appear. This so-called high-resolution scheme should now be oscillation-free near
discontinuities.

In summary, instead of using uR = ui+1 and uL = ui at the cell interface


i + 21 , it is used uR = u+
i+ 12
and uL = u−
i+ 12
, as defined in (A.44)-(A.45), so that the
high-resolution reconstructed jump at the interface i + 12 is given by:

∆uhr
i+ 1
≡ u+
i+ 1
− u−
i+ 1
(A.54)
2 2 2

In other words, the explicit numerical flux was first order :

1  |Jei+ 1 |
1st
fi+ 1 = f(ui+1 ) + f(ui ) − 2
· ∆u1st
i+ 12
(A.55)
2 2 2

And it is now :

1 +  |Jei+ 1 |
hr
fi+ 1 = f(ui+ 1 ) + f(u−
i+ 21
) − 2
· ∆uhr
i+ 12
(A.56)
2 2 2 2

n
A. Implementation of Up-wind Roe Scheme

A.3.2. Preconditioning

Preconditioning is a procedure used to deal with the typical stiffness prob-


lems [144, 139] encountered when a compressible solver like Roe’s scheme is
used to solve the governing equations of a flow in the low Mach number range.

Defining as usual the Mach number M as the ratio of flow velocity u to sound
speed c, the low Mach number region is roughly defined as M ≤ 0.2.

In two-phase flows, liquid phases are likely to be in the low Mach number region.
Indeed in liquids the sound speed is generally by far greater than the flow velocity.
Similarly, the vapor phase is generally outside this low Mach number range, unless
there is possible reverse flow phenomena. Moreover, strong variations of the Mach
number are likely to occur at phase transitions, or when the cross-section of the
component is not constant.

Therefore the compressible Roe scheme, suited for M ≥ 0.2 approximatively,


must be modified accordingly as regions with M ≤ 0.2 and M ≥ 0.2 occur si-
multaneously. Since the magnitude of the sound speed relative to that of the
flow velocity is responsible for the stiffness of the compressible flow equations
at low Mach numbers, this problem is dealt with by artificially scaling down
the amplitude of the acoustic waves in order to improve the conditioning of the
system : this is know as preconditioning.

This can be achieved by some algebraic manipulation of the time derivative


terms and the Roe-matrix of the original system of equations. A better condi-
tioning of the system leads to improved accuracy and convergence in steady and
unsteady computations.

The stiffness problem is solved by multiplying the time-derivative of the sys-


tem of equations (A.3) by the preconditioning matrix P −1 :

∂u ∂f
P −1 + =0 (A.57)
∂t ∂x

o
A. Implementation of Up-wind Roe Scheme

In that way, the convergence to steady-sate is accelerated, but the time consistency
is lost. An alternative formulation, performing a correct scaling of the artificial
dissipation terms, but keeping the original form (A.3) of the equations, enables a
time-accurate unsteady computation: rather than using (A.13) for the numerical
flux, it is used:

1  1
fpre (uL , uR ) = f(uL ) + f(uR ) − Pe−1 · |Pe · J(
e uL , uR )| · ∆u (A.58)
2 2

The form of that preconditioning matrix Pe is dependent on the model considered,


and therefore the reader is referred to the corresponding sections of each model,
where its explicit formulation is given. A deeper analysis of preconditioning
methods can be found, for instance, in [139, 81].

A.3.3. Variable cross-section

When the cross-section of the component varies radially along the longitudinal
axis, the conservation laws (A.1) or (A.2) must be modified. A new set of conser-
vative variables is used : uA = A(ρ, ρu, ρE), and the governing set of equations is
:

∂ρA ∂p ∂ρuA
+ Aκw ρ + = 0
∂t ∂t ∂x
 ∂ρ A ∂ρ uA 
nc nc
+ = 0 if needed
∂t ∂x
∂ρuA ∂(ρu2 + p)A dA
+ = p − AFw − Aρg (A.59)
∂t ∂x dx
∂ρEA ∂ρuHA
+ = AQw
∂t ∂x

In order to keep these equations under conservative form, an extra source term
appears in the momentum equation, explicitly showing the influence of the cross-
section variation.

Apart from this extra source term, only a matter of notations distinguishes this set
of equations from the general one (constant cross-section) : the equations remain
unchanged. Consequently, only the constant cross-section case is described in
the next chapters, unless specified otherwise. One has only to remind that an

p
A. Implementation of Up-wind Roe Scheme

extra term should be added if the cross-section were to vary along the axis of the
component.

A way to deal with variable cross-section in EcosimPro could be the following :

• Implement the above formalism to the component continuous block (most


likely the pipe);

• Build another component inherited from the first one, where the change of
cross-section is defined : either the change is zero (constant cross-section) or
the area varies with the longitudinal direction following a given expression
(the smoother the better).

q
B. Friction Factor Correlations

B.1. Single-Phase Friction Factor Calculation. Function


hdc_fric

The function hdc_fric incorporates the evolution of the friction factor as a


function of the local Reynolds number (ρvD/µ) and the roughness ε.
The friction factor (f ) is calculated by means of a simple correlation valid for
laminar, turbulent and transient flow.
" 12 # 121
8 1
f =8· + (B.1)
Re (A + B)3/2

where
 16
1
A = 2.457 ln
(7/Re)0.9 + 0.27ε/D
 16
37530
B=
Re

B.2. Two-Phase Friction Factor Calculation. Friedel


Correlation

The following formulation is taken from reference [135].


The correlation method of Friedel [48] (1979) utilizes a two-phase multiplier:

∆Pf rict = ∆Pl Φ2f r (B.2)

r
B. Friction Factor Correlations

where ∆Pl is calculated for the liquid-phase flow as:


   
L 2 1
∆Pl = 4fl G (B.3)
D 2ρl

The liquid friction factor fl and liquid Reynolds number are obtained from

0.079 GD
f= Re = (B.4)
Re0.25 µ

Using the liquid dynamic viscosity µl . His two-phase multiplier is

3.24 F H
Φ2f r = E + 0.045 We0.035 (B.5)
FrH l

The dimensionless factors FrH , E , F and H are as follows:

G2
FrH =
g D ρ2H
ρl fg
E = (1 − x)2 + x2
ρG fl
F = x0.78 (1 − x)0.224
 0.91  0.19  0.7
ρl µG µg
H= 1−
ρg µL µl

The liquid Weber Wel is defined as

G2 D
Wel = (B.6)
σ ρH

where σ is the surface tension. The following alternative definition of the homo-
geneous density ρH based on vapor quality is used:
 −1
x 1−x
ρH = + (B.7)
ρG ρL

B.3. Elbow Pressure Loss Function

This function calculates the bend pressure drop coefficient. It depends on the
relative radius of curvature, Rbend /D, the relative roughness, ε/D, and the bend

s
B. Friction Factor Correlations

angle, α.
According to Idelchik [63], the total resistance coefficient of pipe bends is the
product of the following coefficients (see Figure B.1):

Figure B.1.: Elbow pressure loss parameters

• angle effect:
r
α α
ξangle = 0.957 + 0.226 + 0.407 sin(α) − 0.833 sin(α/2)
90 90

• radius effect:
 r
R
0.21/ (R/D > 1)


ξradius =  D2.5
R
0.21/ (R/D < 1)


D

• roughness effect:
  2
min(2, 1 + 106 ε )

(R/D > 1.5)
ξroug = D
ε
min(2, 1 + 103
 ) (R/D < 1.5)
D

t
B. Friction Factor Correlations

Then the pressure drop coefficient is:

ξbend = ξangle · ξradius · ξroug (B.8)

u
C. Film Coefficient Calculation

The film coefficient h is evaluated by mean of the Nusselt number calculation.


The correlation used for the Nusselt assessment is function of the quality of the
fluid, therefore there will be a specific correlation for single-phase fluid, two-phase
fluid etc..

Single phase :
Laminar and turbulent Nusselt numbers:

N ulam = 4 (C.1)
N utur = 0.023 Re0.8 P r0.4 (C.2)

The equivalent Nusselt number covering transitions zones is:

N u = (N u16 16 1/16
lam + N utur ) (C.3)

Then, the single phase film coefficient is calculated as follows:


 
λ
hsp = N u (C.4)
D

Condensation (two-phase or vapour and Tw < Tsat ) :


This method is based on Boyko & Kruzhilin’s [37] correlation, appropriate for
film-wise condensation in uniform channels under forced convection conditions:
q
hcond = hsp 1 + x(ρl /ρg − 1) (C.5)

Superheated Condensation (Quality =1 and Tw < Tsat ) :


The method used for the calculation of the heat transfer coefficient is shown below.

v
C. Film Coefficient Calculation

The method was chosen in order to keep a continuity between single and two
phase regimes:
hg (Tsat − Tw ) + hcond (T − Tsat )
h= (C.6)
(T − Tw )
where hg is the vapour single phase film coefficient. hcond is the condensation
correlation film coefficient considering the actual pressure saturation properties.

Boiling (Quality < 0.7 and Tw > Tsat ) :


According to Chen [26], for vapour quality < 0.7 and if the stratification is not
severe, several steps must be followed in order to calculate the film coefficient
under vaporization regime. First of all, the convective film coefficient for liquid
must be calculated:  
0.8 0.4 λl
hl = 0.023 Rel P rl (C.7)
D
where the sub index l makes reference to saturated liquid conditions. The inverse
of the Lockhart-Martinelli [89] parameter is calculated:
 0.9  0.5  0.1
x ρl µg
1/Xtt = (C.8)
(1 − x) ρg µl

Then, the convective boiling contribution is calculated as follows: hc on = F hl ,


where:

2.35 (1/Xtt + 0.213)( 0.736) (1/Xtt > 0.1)
F =
1 (1/Xtt > 0.1)

The nucleate boiling contribution is calculated as follows:

hnuc = B (Tw − Tsat )0.24 (Psat,Tw − Psat )0.75 (C.9)

where

λ0.79
l Cp0.45
l ρ0.49
l S
B = 0.001222 0.29 (C.10)
σ µl (ρg (hg − hl ))0.24
0.5

1
S= Re2ph = Rel , F 1.25 (C.11)
1 + 2.53 exp −6 Re1.172ph

w
C. Film Coefficient Calculation

where σ is the surface tension; hg − hl is the vaporization latent heat. Finally, the
combined boiling film coefficient is hChen = hcon + hnuc

Boiling (Quality > 0.9 and Tw > Tsat ) :


For vapour quality x > 0.9 a post-dry-out correlation due to Dougall & Rohsenow [37]
is used:
 
λg
hg = 0.023 Re0.8 0.4
g P rg (C.12)
D
ρg
Φ = x + (1 − x) (C.13)
ρl
hDR = hg · Φ (C.14)

Boiling (0.7 < Quality < 0.9 and Tw > Tsat ) :


For vapour quality 0.7 < x < 0.9, cubic spline interpolation is performed between
the Chen & Dougall - Rohsenow correlations.

Subcooled Boiling (Quality = 0 and Tw > Tsat ) :


The method used is to calculate the heat transfer coefficient as to ensure continuity
between the single and two phase regimes:

hl (Tsat − Tl ) + hChen (Tw − Tsat )


h= (C.15)
(Tw − T )

where hl is the liquid single phase film coefficient. hChen is the Chen correlation
film coefficient considering the actual pressure saturation properties.

x
Bibliography

[1] R. Abgrall. How to Prevent Pressure Oscillations in Multicomponent Flow


Calculations: a Quasi Conservative Approach. J. Comput. Phys., 14:150–160,
1996.

[2] A. Adamkowski and M. Lewandowski. Experimental Examination of Un-


steady Friction Models for Transient Pipe Flow Simulation. Journal of
Fluids Engineering, 128:1351 – 1363, November 2006.

[3] G. Albano, J. Hebrard, and V. Leudiere. HM7B Engine Transient Simulator


with CARINS tool. In EUCASS, 2005.

[4] M. A. Arguello. The Concept Design of a Split Flow Liquid Hydrogen


Turbopump. Master’s thesis, Air Force Institute of Technology, March 2008.

[5] Astrium and TU Dresden. Final report on data correlation and evaluation
of test results phase 1 part 1. Internal ESA contractor report GSTP-2-TN-06-
Astrium, Astrium, 2001.

[6] M. Atsumi, A. Ogawara, K. Akazawa, and K. Takeishi. Development of


Visual Integrated Simulator for Rocket Engine Cycle. In 4th International
Conference on Launcher Technology, Space Launcher Liquid Propulsion,
Belgium, 2002. CNES.

[7] T. J. Avampato and C. Saltiel. Dynamic Modeling of Starting Capabilities


of Liquid Propellant Rocket Engines. Journal of Propulsion and Power,
Vol.11(No.2):pp.292 – 300, March - April 1995.

[8] F. Azevedo. Propriétés thermodynamiques des ergols. PhD thesis, CNES.

i
Bibliography

[9] D. R. Bartz. Turbulent boundary-layer heat transfer from rapidly acceler-


ating flow of rocket combustion gases and of heated air. Technical Report
NASA-CR-62615, Jet Propulsion Laboratory, 1963.

[10] E.N. Belyaev, V.K. Chvanov, and V.V. Chervak. Mathematical Modelling
of the Workflow of Liquid Rocket Engines. MAI, 1999. (In Russian).

[11] R. Biggs. History of Liquid Rocket Engine Development in the United


States, 1995 - 1980, volume Vol. 13. American Astronautical Society, 2001.

[12] R. Biggs. Space Shuttle Main Engine: the First Twenty Years and Beyond,
volume Vol. 29. American Astornautical Society, 2008.

[13] M. Binder. An RL10A-3-3A Rocket Engine Model Using the Rocket Engine
Transient Simulator (ROCETS) Software. Contractor Report NASA CR-
190786, NASA, July 1993.

[14] M. Binder. A Transient Model of the RL10A-3-3A Rocket Engine. Contractor


Report NASA CR-195478, NASA, July 1995.

[15] M. Binder, T. Tomsik, and J.P. Veres. RL10A-3-3A Rocket Engine Modeling
Project. Technical Memorandum NASA TM-107318, NASA, 1997.

[16] M. P. Binder and J. L. Felder. Predicted Performance of an Integrated


Modular Engine System. In AIAA/SAE/ASME/ASEE 29th Joint Propulsion
Conference and Exhibit, number AIAA 1993-1888, 1993.

[17] J. E. Bradford, A. Charania, and B. St. Germain. REDTOP-2: Rocket Engine


Design Tool Featuring Engine Performance, Weight, Cost, and Reliability. In
40th AIAA/ASME/SAE/ASEE Joint Propulsion Conference and Exhibit,
number AIAA 2004-3514, 2004.

[18] K.E. Brenan, S.L. Campbell, and L.R. Petzold. Numerical Solution of Initial-
Value Problems in Differential-Algebraic Equations. SIAM, 1996.

[19] C. D. Brown. Conceptual Investigations for a Methane-Fueled Expander


Rocket Engine. In 40th AIAA/ASME/SAE/ASEE Joint Propulsion Confer-
ence and Exhibit, number AIAA 2004-4210, 2004.

ii
Bibliography

[20] J. R. Brown. Cryogenic Upper Stage Propulsion - RL10 and Derivative


Engines. Technical report, Pratt&Whitney, 1990.

[21] J. R. Brown, R. R. Foust, D. E. Galler, P. G. Kanic, T. D. Kmiec, C. D. Limerick,


R. J. Peckham, and T. Swartwout. Design and Analysis Report for the RL10-
2B Breadboard Low Thrust Engine. Contract Report NASA-CR-174857,
NASA, 1984.

[22] B. Campbell and R. Davis. Quasi-2D Unsteady Flow Solver Mod-


ule for Rocket Engine and Propulsion System Simulations. In 42nd
AIAA/ASME/SAE/ASEE Joint Propulsion Conference and Exhibit, num-
ber AIAA 2006-5061, 2006.

[23] Van P. Carey. Liquid-Vapor Phase-Change Phenomena. Taylor and Fran-


cis, 1992.

[24] J. Carlile and R. Quentmeyer. An Experimental Investigation of High-


Aspect-Ratio Cooling Passages. In AIAA/SAE/ASME/ASEE 28th Joint
Propulsion Conference and Exhibit, number AIAA 1992-3154, July 1992.

[25] H. Chaudhry. Applied Hydraulic Transients - 2nd Edition. Van Nostrand


Reinhold Company, New York, 1987.

[26] J. C. Chen. A correlation for boiling heat transfer to saturated fluid in


convective flow. I.Eng. Chem. Process Des. Dev., 5:322–329, 1966.

[27] S. Clerc. Numerical Simulation of the Homogeneous Equilibrium Model


for Two-Phase Flows. Journal of Computational Physics, Vol. 161:pp. 354
– 375, 2000.

[28] Kenneth P. Coffin and Cleveland O’Neal, Jr. Experimental thermal con-
ductivities of the N2O4 - 2NO2 system. Technical Note TN-4209, NASA,
1958.

[29] A. M. Crocker and S. Peery. System Sensitivity Studies of a LOX/Methane


Expander Cycle Rocket Engine. In 34th AIAA/ASME/SAE/ASEE Joint
Propulsion Conference and Exhibit, number AIAA 1998-3674, 1998.

iii
Bibliography

[30] A. Dalbies, G. De Spiegeleer, C. Maquet, and E. Robert. Engines Trade Off for
Ariane 2010 Initiative. In AIAA/ASME/SAE/ASEE 38th Joint Propulsion
Conference and Exhibit, number AIAA 2002-3839, 2002.

[31] R. Davis and B. T. Campbell. Quasi-One-Dimensional Unsteady-Flow


Procedure for Real Fluids. AIAA Journal, Vol.45(10):pp.2422 – 2428, October
2007.

[32] M. De Rosa, J. Steelant, and J. Moral. ESPSS: European Space Propulsion


System Simulation. In Space Propulsion Conference, May 2008.

[33] J.C. DeLise and M.H.N. Naraghi. Comparative Studies of Convective Heat
Transfer Models for Rocket Engines. In AIAA/ASME/SAE/ASEE 31st Joint
Propulsion Conference and Exhibit, number AIAA 95-2499, July 1995.

[34] F. Di Matteo and M. De Rosa. Object Oriented Steady State Analysis and
Design of Liquid Rocket Engine Cycles. In 3AF, editor, Space Propulsion
Conference 2010, San Sebastian, Spain, May 2010.

[35] F. Di Matteo, M. De Rosa, and M. Onofri. Semi-Empirical Heat Transfer


Correlations in Combustion Chambers for Transient System Modelling. In
3AF, editor, Space Propulsion Conference 2010, San Sebastian, Spain, May
2010.

[36] F. Di Matteo, M. De Rosa, M. Pizzarelli, and M. Onofri. Modelling of


Stratification in Cooling Channels and its Implementation in a Transient
System Analysis Tool. In AIAA, editor, AIAA/SAE/ASME/ASEE 46th Joint
Propulsion Conference and Exhibit, Nashville, TN, USA, July 2010. AIAA.

[37] R.S. Dougall and W.H. Rohsenow. Film boiling on the inside of the inside of
the vertical tubes with upward flow of the fluid at low qualities. Technical
Report 9079.26, M.I.T., 1963.

[38] S. Durteste. A Transient Model of the VINCI Cryogenic Upper Stage Rocket
Engine. In 43rd AIAA/ASME/SAE/ASEE Joint Propulsion Conference
and Exhibit, number AIAA 2007-5531, 2007.

iv
Bibliography

[39] Y. Elkouch. Physical Model Specification Cenaero’s Contribution. Technical


Note TN-3221, CENAERO, 2007.

[40] Empresarios Agrupados. Ecosimpro: Continuous and discrete modelling


simulation software. http://www.ecosimpro.com, 2007.

[41] Empresarios Agrupados. ESPSS-2 industrial evaluation. Internal ESA


document, May 2009.

[42] Empresarios Agrupados. ESPSS user manual. 1.4.1 edition, 2009.

[43] Empresarios Agrupados. ESPSS user manual. 2.0 edition, 2010.

[44] Capcom Espace. Le moteur HM7B. http://www.capcomespace.net


/dossiers/espace_europeen/ariane/ariane4/moteur_HM7B.htm. retrieved
28/04/2010.

[45] Capcom Espace. Le moteur Vinci. http://www.capcomespace.net


/dossiers/espace_europeen/ariane/. retrieved 28/10/2011.

[46] Capcom Espace. Le moteur Vulcain 2. http://www.capcomespace.net


/dossiers/espace_europeen/ariane/. retrieved 28/10/2011.

[47] A. Espinosa Ramos. CARAMEL: The CNES computation sotware for design-
ing Liquid ROcket Engines. In 4th International Conference on Launcher
Technology, Space Launcher Liquid Propulsion, 2002.

[48] L. Friedel. Improved Friction Pressure Drop Correlations for Horizontal


and Vertical Two Phase Pipe Flow. In European Two Phase Flow Group
Meeting, Italy, 1979. Ispra.

[49] A. Frohlich, M. Popp, G. Schmidt, and D. Thelemann. Heat Transfer Char-


acteristics of H2/O2 - Combustion Chambers. In AIAA/SAE/ASME/ASEE
29th Joint Propulsion Conference, number AIAA 93-1826, June 1993.

[50] Y. Fukushima, H. Naktsuzi, R. Nagao, K. Kishimoto, K. Hasegawa, T. Ko-


ganezawa, and S. Warashina. Development Status of LE-7A and LE-5B
Engines for H-IIA Family. Acta Astronautica, Vol. 50:pp. 275 – 284, 2002.

v
Bibliography

[51] F. Gao, Y. Chen, and Z. Zhang. Numerical Simulation of Steady and


Filling Process of Low Temperature Liquid Propellants Pipeline. In Fifth
International Conference on Fluid Mechanics. Tsinghua University Press
& Springer, August 2007.

[52] A. J. Glassman. Computer Code for Preliminary Sizing Analysis of Axial-


Flow Turbines. Contractor Report CR-4430, NASA, 1992.

[53] S. Go. A historical survey with success and maturity estimates of launch
systems with RL10 upper stage engines. In IEEE, editor, Reliability and
Maintainability Symposium, 2008.

[54] S.K. Godunov. A Difference Scheme for Numerical Computation of Dis-


continuous Solutions of Equations of Fluid Dynamics. Math. Sbornik,
47(89):271–306, 1959.

[55] Sanford Gordon and Bonnie J. McBride. Computer program for calculation
of complex chemical equilibrium compositions and applications. Technical
Report RP-1311, NASA, 1994.

[56] T. M. Haarmann. Numerische Simulation des Warmeübergangs in einer


kryogenen Raketenbrennkammer. PhD thesis, RWTH Aachen, 2006.

[57] T. M. Haarmann and W. W. Koschel. Computation of Wall Heat


Fluxes in Cryogenic H2/O2 Rocket Combustion Chambers. In
AIAA/ASME/SAE/ASEE 38th Joint Propulsion Conference and Exhibit,
number AIAA 2002-3693, July 2002.

[58] M. S. Haberbusch, C. T. Nguyen, and A. F. Skaff. Modeling RL10


Thrust Increase with Densified LH2 and LOX Propellants. In 39th
AIAA/ASME/SAE/ASEE Joint Propulsion Conference and Exhibit, num-
ber AIAA 2003-4485, July 2003.

[59] M. S. Haberbusch, A. F. Skaff, and C. T. Nguyen. Modeling the RL-


10 with Densified Liquid Hydrogen and Oxygen Propellants. In 38th
AIAA/ASME/SAE/ASEE Joint Propulsion Conference & Exhibit, number
AIAA 2002-3597, July 2002.

vi
Bibliography

[60] H. Hearn. Development and Application of a Priming Surge Analysis


Methodology. In 41st AIAA/ASME/SAE/ASEE Joint Propulsion Confer-
ence and Exhibit, number AIAA 2005-3738, 2005.

[61] W. D. Huang. An Object-Oriented Analysis Method for Liquid Rocket


Engine System. In 36th AIAA/ASME/SAE/ASEE Joint Propulsion Confer-
ence and Exhibit, number AIAA 2000-3770, 2000.

[62] Dieter K. Huzel and David H. Huang. Modern Engineering for Design of
Liquid-Propellant Rocket Engines, volume 147 of Progress in Astronautics
and Aeronautics. AIAA, 1992.

[63] I. E. Idelchik. Handbook of Hydraulic Resistance. Begell House, 3rd edition,


2001.

[64] Inc.. Information Systems Laboratories. RELAP5/MOD3.3 CODE MAN-


UAL VOLUME IV: MODELS AND CORRELATIONS. Nuclear Safety
Analysis Division, Rockville, Maryland, December 2001.

[65] A. Isselhorst. HM7B Simulation with ESPSS Tool on ARIANE 5 ESC-A Up-
per Stage. In 46th AIAA/ASME/SAE/ASEE Joint Propulsion Conference
and Exhibit, number AIAA 2010-7047, 2010.

[66] K. J. Kacynski. Thermal stratification potential in rocket engine coolant


channels. Technical Memorandum NASA-TM-4378, NASA, 1992.

[67] G. P. Kalmykov, E. V. Lebedinsky, V. I. Tararyshkin, and I. O. Yeliseev.


Expander LRE of 200 tf thrust on hydrocarbon fuel. In 4th International
Conference on Launcher Technology, Space Launcher Liquid Propulsion,
2002.

[68] V. M. Kalnin and V. A. Sherstiannikov. Hydrodynamic Modelling of the


Starting Process in Liquid-Propellant Engines. Acta Astronautica, Vol. 8:pp.
231 – 242, 1981.

[69] V. M. Kalnin and V. A. Sherstiannikov. Vibration and Pulsation Processes in


feed Systems of Liquid Rocket Engines. Acta Astronautica, Vol. 10:pp. 713
– 718, 1983.

vii
Bibliography

[70] A. Kanmuri, T. Kanda, Y. Wakamatsu, Y. Torri, E. Kagawa, and


K. Hasegawa. Transient Anaysis of LOX/LH2 Rocket Engine (LE-7) . In
25th AIAA/ASME/SAE/ASEE Joint Propulsion Conference, number AIAA
89-2736, July 1989.

[71] H. Karimi and R. Mohammadi. Modeling and simulation of a two com-


bustion chambers liquid propellant engine. Aircraft Engineering and
Aerospace Technology, Vol.79:pp. 390 – 397, 2007.

[72] H. Karimi, R. Mohammadi, and E. Taheri. Dynamic Simulation and Paramet-


ric Study of a Liquid Propellant Engine. In 3rd International Conference
on Recent Advances in Space Technologies, 2007.

[73] H. Karimi, A. Nassirharand, and M. Beheshti. Dynamic and Nonlinear


Simulation of Liquid-Propellant Engines. Journal of Propulsion and Power,
Vol.19(No.5):pp. 938 – 944, September - October 2003.

[74] J. Keppeler, E. Boronine, and F. Fassl. Startup Simulation of Upper Stage


Propulsion System of ARIANE 5. In 4th International Conference on
Launcher Technology, Space Launcher Liquid Propulsion, 2002.

[75] T. Kimura, M. Sato, T. Masuoka, T. Kanda, and A. Osada. Effects of Deep


Throttling on Rocket Engine Systems . In 46th AIAA/ASME/SAE/ASEE
Joint Propulsion Conference and Exhibit, number AIAA 2010-6727, 2010.

[76] A. E. Krach and A. M. Sutton. Another Look at the Practical and


Theoretical Limits of an Expander Cycle, LOX/H2 Engine. In 35th
AIAA/ASME/SAE/ASEE Joint Propulsion Conference and Exhibit, 1999.

[77] L. Kun and Z. Yulin. A Study on Versatile Simulation of Liquid Propellant


Rocket Engine Systems Transients. In AIAA/ASME/SAE/ASEE 36th Joint
Propulsion Conference and Exhibit, number AIAA 2000-3771, July 2000.

[78] J. C. Leahy, S. G. Hanna, J. Y. Malchi, and E. S. Kim. Liquid Propulsion:


Engine Production and Operation. In Encyclopedia of Aerospace Enngi-
neering. John Wiley & Sons, 2010.

viii
Bibliography

[79] F. LeBail and M. Popp. Numerical Analysis of High Aspect Ratio Cooling
Passage Flow and Heat Transfer. In AIAA/SAE/ASME/ASEE 29th Joint
Propulsion Conference, number AIAA 1993-1829, June 1993.

[80] A. LeClair and A. K. Majumdar. Computational Model of the Chill-


down and Propellant Loading of the Space Shuttle External Tank. In
AIAA/ASME/SAE/ASEE 46th Joint Propulsion Conference and Exhibit,
number AIAA 2010-6561, July 2010.

[81] D. Lee. Local Preconditioning of the Euler and Navier-Stokes Equations.


PhD thesis, University of Michigan, 1996.

[82] B. Legrand, G. Albano, and P. Vuillermoz. Start up transient modelling


of pressurised tank engine: AESTUS application. In 4th International
Conference on Launcher Technology, Space Launcher Liquid Propulsion,
2002.

[83] A. J. Letson and C. Christian. RL10A-3-3 Rocket Engine Oxidizer Pump


Development Program. Contract Report NASA-CR-83547, NASA, 1966.

[84] V. Leudiére, P. Supié, and M. Villa. KVD1 Engine in LOX/CH4. In


AIAA/ASME/SAE/ASEE 43rd Joint Propulsion Conference & Exhibit,
number AIAA 2007-5446, 2007.

[85] R.J. Leveque. Finite Volume Methods for Hyperbolic Problems. Cambridge
Texts in Applied Mathematics, United Kingdom, 2002.

[86] LewisResearchCenter. Centaur Space Vehicle Pressurized Propellant Feed


System Tests. Technical Note TN-D-6876, NASA, 1972.

[87] T. Y. Lin and D. Baker. Analysis and Testing of Propellant Feed System
Priming Process. Journal of Propulsion and Power, Vol. 11(No.3):pp. 505 –
512, May - June 1995.

[88] P.J. Linstrom and W.G. Mallard. NIST Chemistry WebBook, NIST Stan-
dard Reference Database Number 69. National Institute of Standards and
Technology, Gaithersburg MD, 20899, 2011. http://webbook.nist.gov, (re-
trieved July 14, 2011).

ix
Bibliography

[89] R. W. Lockhart and R. C. Martinelli. Proposed Correlation of Data for


Isothermal Two Phase Flow, Two Component Flow in Pipes. Chem. Eng.
Prog., Vol. 45:pp. 38 – 49, 1949.

[90] P. C. Lozano-Tovar. Dynamic Models for Liquid Rocket Engines with


Health Monitoring Application. Master’s thesis, Massachusetts Institute of
Technology, 1998.

[91] C. Manfletti. Transient Behaviour Modelling of Liquid Rocket Engine


Components. PhD thesis, Maschinenwesen der Rheinisch-Westfälischen
Technischen Hochschule Aachen, 2009.

[92] C. Manfletti. Start-Up Transient Simulation of a Pressure Fed LOX/LH2


Upper Stage Engine Using the Lumped Parameter-based MOLIERE Code. In
46th AIAA/ASME/SAE/ASEE Joint Propulsion Conference and Exhibit,
number AIAA 2010-7046, 2010.

[93] C. Manfletti. Water Hammer Simulation Using a Lumped Parameter Model.


In IAF, editor, Space Propulsion Conference, 2010.

[94] D. Manski, C. Goertz, H. Saßnick, J. R. Hulka, B. D. Goracke, and D. J. H.


Levack. Cycles for Earth-to-Orbit Propulsion. Journal of Propulsion and
Power, Vol. 14(No.5):pp. 588–604, September-October 1998.

[95] N. B. McNelis and M. S. Haberbusch. Hot Fire Ignition Test with Densified
Liquid Hydrogen Using a RL10B-2 Cryogenic H2/O2 Rocket engine. In
33rd AIAA/ASME/SAE/ASEE Joint Propulsion Conference and Exhibit,
number AIAA 1997-2688, 1997.

[96] M. Meyer. Electrically Heated Tube Investigation of Cooling Channel Geom-


etry Effects. In 31st ASME, SAE, and ASEE, Joint Propulsion Conference
and Exhibit, number AIAA 1995-2500, 1995.

[97] NASA. Liquid Rocket Engine Centrifugal Flow Turbopumps. Technical


Report SP-8109, NASA, December 1973.

[98] NASA. Turbopump Systems for Liquid Rocket Engines. Technical Report
SP-8107, NASA, August 1974.

x
Bibliography

[99] J. Nathman, J. Niehaus, J. C. Sturgis, A. Le, and J. Yi. Preliminary


Study of Heat Transfer Correlation Development and Pressure Loss
Behavior in Curved High Aspect Ratio Coolant Channels. In 44th
AIAA/ASME/SAE/ASEE Joint Propulsion Conference and Exhibit, num-
ber AIAA 2008-5240, 2008.

[100] D. Nguyen and A. Martinez. Versatile Engine Design Software. In 28th


AIAA/ASME/SAE/ASEE Joint Propulsion Conference and Exhibit, num-
ber AIAA 1993-2164, 1993.

[101] G. Odonneau, G. Albano, V. Leuduere, and J. Masse. Carins: A New


Versatile and Flexible Tool for Engine Transient Prediction - development
status. In 6th International Symposium on Launcher Technologies, 2005.

[102] G. Odonneau, G. Albano, and J. Masse. Carins: A New Versatile and Flexible
Tool for Engine Transient Prediction. In 4th International Conference on
Launcher Technology, Space Launcher Liquid Propulsion, 2002.

[103] A. J. Pavli, J. K. Curley, P. A. Masters, and R. M. Schwartz. Design and


Cooling Performance of a Dump Cooled Rocket Engine. Technical Note TN
D-3532, NASA, August 1966.

[104] S. Peery and A. Minick. Design and Development of an Advanced Expander


Combustor. In 34th AIAA/ASME/SAE/ASEE Joint Propulsion Conference
and Exhibit, number AIAA 1998-3675, 1998.

[105] P. Pempie and L. Boccaletto. LOX/CH4 Expander Upper Stage Engine. In


IAF, editor, 55th International Astronautical Congress, 2004.

[106] M. Pizzarelli, F. Nasuti, and M. Onofri. A Simplified Model for the Analysis
of Thermal Stratification in Cooling Channels. In EUCASS, 2nd European
Conference for Aerospace Sciences, July 2007.

[107] M. Pizzarelli, F. Nasuti, and M. Onofri. Flow Analysis of


Transcritical Methane in Rectangular Cooling Channels. In 44th
AIAA/ASME/SAE/ASEE Joint Propulsion Conference and Exhibit, num-
ber AIAA 2008-4556, July 2008.

xi
Bibliography

[108] M. Pizzarelli, F. Nasuti, and M. Onofri. Investigation of Transcriti-


cal Methane Flow and Heat Transfer in Curved Cooling Channels. In
45th AIAA/ASME/SAE/ASEE Joint Propulsion Conference, number AIAA
2009-5304, August 2009.

[109] M. Pizzarelli, F. Nasuti, R. Paciorri, and M. Onofri. Numerical Analysis of


Three-Dimensional Flow of Supercritical Fluid in Asymmetrically Heated
Channels. AIAA Journal, 47(11):pp. 2534–2543, November 2009.

[110] Pratt&Whitney. Design Report For RL10A-3-3 Rocket Engine. Contractor


Report CR-80920, NASA, 1966.

[111] Pratt&Whitney. Rl10a-3-3 rocket engine oxidizer pump development pro-


gram. Contract Report CR-83547, NASA, 1966.

[112] Pratt&Whitney. Design Study of RL-10 Derivatives - Final Report. Technical


Report CR-120145, NASA, 1973.

[113] Pratt&Whitney. Design and Analysis Report for the RL10-IIB Breadboard
Low Thrust Engine. Contract Report NASA-CR-174857, NASA, 1984.

[114] Pratt&Whitney. RL10 ignition limits test for Shuttle Centaur. Contract
Report NASA-CR-183199, NASA, 1987.

[115] Pratt&Whitney. Shuttle Centaur Engine Cooldown Evaluation and Effects


of Expanded Inlets on Start Transient. Contract Report NASA-CR-183198,
NASA, 1987.

[116] D. Preclik, D. Wiedmann, W. Oechslein, and J. Kretschmer. Cryogenic


Calorimeter Chamber Experiments and Heat Transfer Simulations. In
AIAA/ASME/SAE/ASEE 34th Joint Propulsion Conference and Exhibit,
number AIAA 1998-3440, July 1998.

[117] D. C. Pytanowski. Rocket-Engine Control-System Reliability-Enhancement


Analysis. In IEEE, editor, Annual RELIABILITY and MAINTAINABIL-
ITY Symposium, 1999.

xii
Bibliography

[118] V. Rachuk and N. Titkov. The First Russian LOX-LH2 Expander Cycle LRE:
RD0146. In 42nd AIAA/ASME/SAE/ASEE Joint Propulsion Conference,
number AIAA 2006-4904, July 2006.

[119] D. Ramesh and M. Aminpoor. Nonlinear, Dynamic Simulation of an


Open Cycle Liquid Rocket Engine. In 43rd AIAA/ASME/SAE/ASEE Joint
Propulsion Conference and Exhibit, number AIAA 2007-5507, 2007.

[120] R. Reid, J.M Prausnitz, and B. Poling. The Properties of Gases and Liquids.
McGraw-Hill, 1987.

[121] R. Rhote-Vaney, V. Thomas, and A. Lekeux. Transient Modeling of Cryo-


genic Rocket Engines a Modular Approach. In 4th International Confer-
ence on Launcher Technology, Space Launcher Liquid Propulsion, 2002.

[122] José Ramón Alarcón Rodríguez. LOX/H2 engine heat transfer, equilibrium
and heat transfer analysis program. Technical report, ESA-ESTEC, TEC-
MPC, 2001. Draft version.

[123] P.L. Roe. Approximate Riemann solvers, parameter vectors, and difference
schemes. J. Comput. Phys., 43:357–372, 1981.

[124] E. K. Ruth, H. Ahn, R. L. Baker, and M. A. Brosmer. Advanced Liquid


Rocket Engine Transient Model. In AIAA/ASME/SAE/ASEE 26th Joint
Propulsion Conference and Exhibit, 1990.

[125] G. Saßnick, H. D.and Krülle. Numerical simulation of transients in feed


systems for cryogenic rocket engines . In ASME, SAE, and ASEE, Joint
Propulsion Conference and Exhibit, 31st, number AIAA 1995-2967, 1995.

[126] G. Schmidt, M. Popp, and T. Fröhlich. Design Studies for a 10 Ton Class High
Performance Expander Cycle Engine. In 34th AIAA/ASME/SAE/ASEE
Joint Propulsion Conference and Exhibit, number AIAA 1998-3673, 1998.

[127] R. Schuff, M. Maier, O. Sindiy, C. Ulrich, and S. Fugger. Integrated Modeling


& Analysis for a LOX/Methane Expander Cycle Engine Focusing on Re-
generative Cooling Jacket Design. In 42nd AIAA/ASME/SAE/ASEE Joint
Propulsion Conference, number AIAA 2006-4534, July 2006.

xiii
Bibliography

[128] P. F. Seitz and R. F. Searle. Space Shuttle Main Engine Control System.
In National Aerospace Engineering and Manufacturing Meeting, number
Paper 740927. Society of Automotive Engineers, October 1973.

[129] R. Sekita, A. Watanabel, K. Hirata, and T. Imoto. Lessons Learned from H-2
Failure and Enhancement of H-2A Project. Acta Astronautica, Vol. 48:pp.
431 – 438, 2001.

[130] A. J. Stepanoff. Centrifugal and Axial Flow Pumps, 2nd edition. J. Wiley
and Sons, 1957.

[131] D. Suslov and M. Oschwald. Testfälle zur Validierung von Software für die
Berechnung des Wärmeübergangs in regeneraativ gekühlten Schbkammern.
Technical Report DLR-LA - WT-DO-019, DLR, 2010.

[132] D. Suslov, A. Woschnak, J. Sender, and M. Oschwald. Test specimen design


and measurement technique for investigation of heat transfer processes
in cooling channels of rocket engines under real thermal conditions. In
AIAA/ASME/SAE/ASEE 39th Joint Propulsion Conference and Exhibit,
number AIAA 2003-4613, July 2003.

[133] George Paul Sutton and Oscar Biblarz. Rocket propulsion elements. John
Wiley & Sons, New York, NY, USA, 7th edition, 2001.

[134] A. Tarafder and S. Sarangi. CRESP-LP - A Dynamic Simulator for Liquid-


Propellant Rocket Engines. In 36th AIAA/ASME/SAE/ASEE Joint Propul-
sion Conference and Exhibit, number AIAA 2000-3768, 2000.

[135] J. R. Thome. Engineering Data Book III. Wolverine Tube. Inc, 2010.

[136] A. P. Tishin and L. P. Gurova. Liquid Rocket Engine Modeling. Proceedings


of the Aircraft Engineering College, Vol. 32(3):pp. 99 – 101, 1989.

[137] T. M. Tomsik. A Hydrogen-Oxygen Rocket Engine Coolant Passage Design


Program (RECOP) for Fluid-cooled Thrust Chambers and Nozzles. Technical
Note NASA-N95-70893, NASA, 1994.

xiv
Bibliography

[138] E. F. Toro. Riemann Solvers and Numerical Methods for Fluid Dynamics:
a Practical Introduction. Springer, 1999.

[139] E. Turkel. Preconditioned methods for solving the incompressible and low
speed compressible equations. J. Comput. Phys., 72:277–298, 1987.

[140] A. Vaidyanathan, J. Gustavsoon, and C. Segal. One- and Three-Dimensional


Wall Heat Flux Calculations in a O2/H2 System. Journal of Propulsion and
Power, Vol. 26(No.1):pp. 186 – 188, January - February 2010.

[141] W. M. Van Lerberghe, J. L. Emdee, and R. R. Foust. Enhanced Reliability


Features of the RL10E-1 Engine. Acta Astronautica, Vol. 41(No.4):pp. 197 –
207, 1997.

[142] J. P. Veres. Centrifugal and Axial Pump Design and Off-Design Performance
Prediction. Technical Memorandum NASA-TM-106745, NASA, 1994.

[143] J. P. Veres and T. M. Lavelle. Mean Line Pump Flow Model in Rocket Engine
System Simulation. Technical Memorandum NASA-TM-2000-210574, NASA,
November 2000.

[144] G. Volpe. Performance of Compressible Flow Codes at Low Mach Number.


AIAA, 31(1):49–56, 1993.

[145] M. F. Wadel. Comparison of High Aspect Ratio Cooling Channels Designs


for a Rocket Combustion Chamber with Development of an Optimized
Design. Technical Memorandum NASA-TM-1998-206313, NASA, January
1998.

[146] M. F. Wadel and M. L. Meyer. Validation of High Aspect Ratio Cooling in a


89 kN (20,000 lb) Thrust Combustion Chamber. Technical Memorandum
NASA-TM-107270, NASA, 1996.

[147] G. B. Wallis. One-Dimensional Two Phase Flow. McGraw-Hill, 1969.

[148] D. W. Way and J. R. Olds. SCORES: Developing an Object-Oriented Rocket


Propulsion Analysis Tool. In 34th AIAA/ASME/SAE/ASEE Joint Propul-
sion Conference & Exhibit, number AIAA 1998-3227, 1998.

xv
Bibliography

[149] Wikipedia. Space Shuttle Main Engine.


http://en.wikipedia.org/wiki/Space_Shuttle_Main_Engine. retrieved
28/12/2011.

[150] A. Woschnak and M. Oschwald. Thermo- and Fluidmechanical Analysis


of High Aspect Ratio Cooling Channels. In AIAA/ASME/SAE/ASEE 37th
Joint Propulsion Conference and Exhibit, number AIAA 2001-3404, July
2001.

[151] A. Woschnak, D. Suslov, and M. Oschwald. Experimental and Numerical


Investigations of Thermal Stratification Effects. In AIAA/ASME/SAE/ASEE
39th Joint Propulsion Conference and Exhibit, number AIAA 2003-4615,
July 2003.

[152] N. Yamanishi, T. Kimura, M. Takahashi, K. Okita, and H. Negishi. Transient


Analysis of the LE-7A Rocket Engine Using the Rocket Engine Dynamic
Simulator (REDS). In AIAA/ASME/SAE/ASEE 40th Joint Propulsion Con-
ference and Exhibit, number AIAA 2004-3850, July 2004.

[153] Yanzhong, Cui, Chen, and Yuanyuan. Pressure wave propagation character-
istics in a two-phase flow pipeline for liquid-propellant rocket. Aerospace
Science and Technology, 15:453–464, September 2011.

xvi

Das könnte Ihnen auch gefallen