Sei sulla pagina 1di 15

AIAA 2008-952

46th AIAA Aerospace Sciences Meeting and Exhibit


7 - 10 January 2008, Reno, Nevada

Real gas CFD simulation of supercritical H2-LOX


combustion in the Mascotte single-injector combustor using
a commercial CFD code
Maria M. Poschner1 and Michael Pfitzner2
Universtt der Bundeswehr Mnchen, Neubiberg, 85577, Germany

Modern high performance rocket combustion engines operate at very high pressures up
to 20 MPa. Propellants, typically hydrogen and oxygen, are injected at very low
temperatures, below the critical temperature of oxygen (154.581 K), whereas the pressure
has a supercritical value (p > pc = 50.43 bar) so that mixing and combustion occur at
transcritical conditions. The material properties and equation of state in this region differ
significantly from those at low pressures and high temperatures. The paper presents an
implementation of real gas models into the commercial CFD-Code ANSYS CFX. This solver
is validated with experimental data measured on the Mascotte test rig (V03) at ONERA
which have been published in the 2nd International Workshop on Rocket Combustion
Modelling.

Nomenclature
Variables
c
C
D
htot
k
M
p
Pr
R
S
T
ui
V
xi
yi

=
=
=
=
=
=
=
=
=
=
=
=
=
=
=

specific heat capacity


constant
kinematic diffusivity
total enthalpy
turbulent kinetic energy
molar mass
pressure
Prandtl number
universal gas constant
source term
temperature
velocity component
molar volume
Cartesian coordinates
mole fraction of component i

=
=
=
=
=
=
=
=

Kronecker delta
turbulent eddy dissipation
diffusion coefficient
molecular (dynamic) viscosity
hydrogen-bonding correction factor
thermal conductivity
dimensionless dipole moment
density

Ph. D. student, Thermodynamics Institute LRT-10, Universitt der Bundeswehr, Neubiberg, 85577, Germany,
AIAA Member.
2
Professor, Thermodynamics Institute LRT-10, Universitt der Bundeswehr, Neubiberg, 85577, Germany.
1
American Institute of Aeronautics and Astronautics
092407

Copyright 2008 by the American Institute of Aeronautics and Astronautics, Inc. All rights reserved.


()v

=
=
=
=
=

hard sphere diameter


atomic diffusion volume
stress tensor
integral of potential energy function
Pitzers accentric factor

=
=
=
=
=
=

critical
turbulent
reduced
at low pressure
at constant volume
at constant pressure

Indices
c
t
r
0
v
p

I. Introduction

HE technology of cryogenic rocket combustion engines has been used successfully for many years in several
launchers (Ariane 1 to 5, Space Shuttle, . . . ) and is well known today. Nowadays a new challenge is the
requirement of rising performance and reliability at even more restricted budgets and shortened development cycles.
For this reason the application of computational methods in the development process increases continuously.
Modern high performance rocket combustion engines use cryogenic propellants at high pressures in order to
improve the specific momentum. Typically, cryogenic hydrogen as fuel and liquid oxygen as oxidiser are injected
separately into the combustion chamber, and burn as a diffusion flame. Oxygen is in general in a transcritical
condition because the pressure is above the critical one, although the temperature is subcritical. Under such extreme
conditions the material properties and equation of state of fuel and oxidizer differ significantly from those at
atmospheric pressure.
Above the critical pressure and temperature of the mixture liquid and gaseous phases are no longer separated. On
the liquid side of the critical isotherm the material behaves more like a very dense gas than a liquid. Near the critical
point small changes of state have great effects on transport properties and variables of state, leading to huge
gradients in the density and other thermodynamic variables during the mixing of trans- or supercritical fluids. These
extreme conditions make accurate predictions of the processes taking place very difficult.
Today some very detailed investigations of this mixing process are available. There are studies on the
evaporation, mixing and combustion of liquid or supercritical droplets in a quiescent or moving gaseous surrounding
(Miller et. al. 1998 23; Yang 2000 24; Delplanque and Sirignano 1993 25; Daou et. Al. 1995 26) and supercritical shear
flows as occurring at the coaxial injection into the rocket combustor by DNS (Miller et. al. 2000 21; Oefelein 2006
22
) and by LES (Zong and Yang 2005 27; Oefelein 2006 22). These detailed investigations are very important for the
understanding of the complex mixing and combustion processes occurring in rocket combustion chambers but
nowadays these methods are still too computationally expensive to use them for industrial applications. Here still
RANS methods are applied to support the design process. Two of the latest publications using a RANS method for
the description of a supercritical injection and combustion process are Cheng and Farmer 28 (2006) and Benarous et.
al. 29 (2007). Both validated their non-commercial codes using data of the Mascotte test case RCM-3 presented on
the 2nd IWRCM and were able to reach a very good fit with the experiment. The goal of the present work is to
investigate, how well these results can be reproduced using the commercial CFD-code ANSYS CFX 5.
First a grid converged solution for this test case was generated, using the default equation of state for
supercritical mixing implemented. Two different combustion models were tested. Since by default very simple low
pressure methods are used to represent the material properties in ANSYS the description of fuel and oxidizer was
improved by varying the equation of state and using empirical methods for dynamic viscosity and thermal
conductivity. As hydrogen is a very diffusive material simulations were performed taking differential diffusion
effects into account. Finally effects of the turbulence modelling were investigated.
The probably most serious simplification made in this flow solver is the assumption of ideal mixing. This is
investigated by comparing the solution obtained by the flow solver to thermodynamic data from different real gas
mixing rules.

2
American Institute of Aeronautics and Astronautics
092407

II. Theoretical formulation


A. System of governing equations
In CFX the following system of governing equations is solved. The continuity equation for multi-component
flows is given by:

Yi U jYi

+
=
t
x j
x j

Yi
Yi ''U j '' + Si
i

x j x j

(1)

with

Yi ''U j '' =

t Yi

(2)

Sct x j

Where i is the mass-average density of the component i, is the average mixture density and Yi =  i is the
mass-average mass fraction of component i, U j is the mass-average velocity field, and Si is the source term for
component i accounting for a reaction or combustion model. t is the turbulent viscosity and Sct is the turbulent
Schmidt number which is here identical to the turbulent Prandtl number in order to treat turbulent diffusion
processes for heat and species identically.
The turbulent scalar fluxes Yi ''U j '' are modeled using an eddy dissipation assumption (Eq. 2). Molecular
diffusion caused by temperature gradient (Soret effect), as well as thermal diffusion caused by a concentration
gradient (Dufour effect) are neglected. Here i is the molecular diffusion coefficient which is assumed to be equal
to Di with Di is the kinematic diffusivity, calculated by default in CFX using the bulk viscosity, and modeled in
the investigations for this paper as described below.
In CFX, for constant density the Reynolds averaged equations are solved for momentum and energy:
(ui ) (ui u j )
p

=
+
ij ui/ u /j + S M ,i
+
x j
x j x j
t

2 u

k
ij = i + j
ij
x
x
3
x

j
i
k

( u j htot )
(htot ) p

u /j h +
=
(u j ij ) + S E

t
t

x
x
x

j
j
j
j
j
i xi j

(3)

(4)

(5)

Here is the density and ui are the velocity components. is the stress tensor accounting for the molecular viscosity
of the fluid. In the energy equation htot is the total enthalpy and is the thermal conductivity. SM and SE are source
terms in momentum and energy. Radiation is neglected in the investigations performed for this paper.
The Reynolds flux and stress terms u /j h and ui/ u /j are modeled using the eddy diffusivity hypothesis
( u / h = h ) and the eddy viscosity hypothesis ( u / u / = ui + u j 2 u k 2 k ). For the turbulent
j
t
ij
ij
i j
t
x j
x j xi 3 xk 3
diffusion factor t is modelled by t = t Prt with the turbulent Prandtl number Prt, which is a model constant (CFX
default = 0.9). The system can be closed now by choosing a model for the turbulent viscosity t. This is done in this
paper by a simple two-equation turbulence model, solving two additional transport equations for the turbulent
kinetic energy k and eddy dissipation . It applies t = (C k 2 ) .5
For non-constant density flows and combustion, the Reynolds averaging is replaced by mass (Favre) averaging.
3
American Institute of Aeronautics and Astronautics
092407

B. Property evaluation
Density
The above system of equations has to be closed by an expression giving a relation between Temperature T,
pressure p and the molar volume V. For low pressures this relation is given approximately by the ideal equation of
state (Eq. 6).
pV = RT

(6)

With the universal gas constant R = 8.314472 [J/mol K].


At high pressures and low temperatures this relation differs significantly from that of the ideal gas equation. In
the investigations for this paper two different real gas equations of state were used. One is a two parameter equation,
the Redlich-Kwong equation
p=

RT
a

(V b)
TV (V + b)

(7)

With a = 0.42748 R 2Tc2.5 pc and b = 0.08664 RTc pc . The index c indicates critical properties.
The second real gas equation used is the Peng-Robinson equation. This equation is a three parameter equation
accounting additionally for the deviation of the molecule from spherical shape. This is realized using Pitzers
acentric factor .
p=

RT
a (T )

(V b) V 2 + 2Vb b 2

(7)

With a = 0.457235 R 2Tc2 pc , b = 0.077796 RTc pc and = (1 + (1 T Tc )) 2 with

= 0.37464 + 1.54226 0.26992 2 .


Also the mixing process is different in low- and high-pressure state. At low pressures mixing by partial volumes
is approximately correct. This law assumes that the volume of a multi-component mixture at a given pressure is
identical to the sum of the volumes of the single components at the same pressure.
At higher pressures molecules interact, and the mixture volume differs from Daltons law. There are many
different models for the mixture density. Two very common models are the mixing by parameters and by critical
properties. In these models, it is assumed that a mixture at a given composition follows the principle of
corresponding states.
For the mixing by parameter in the case of the Peng-Robinson equation a, b and are determined for the
mixture. There are many different mixing rules in use. In this paper the classical mixing rule was applied.

a = yi y j aij

(8)

b = yi bi

(9)

= yii

(10)

4
American Institute of Aeronautics and Astronautics
092407

aij is calculated as aii = ai and aij = ai a j (1 kij ) for i j where kij represent the binary interaction parameters,
which are set to zero, as there were no values available.
When mixing by critical properties21 in the case of the Peng-Robinson the mixing rule of equation 8 is used, but
the evaluation of the aij is different:
aij = a (Tc ( i , j ) , pc ,( i , j ) ) (Tr (i , j ) )

(11)

(Tr (i , j ) ) = (1 + (ij )(1 Tr ( i , j ) )) 2

(12)

a (Tc ( i , j ) , pc ,( i , j ) ) = 0.457236 R 2Tc2(i , j ) pc ,( i , j )

(13).

(ij ) = 0.379642 + 1.48503ij 0.164423ij2 + 0.016666ij3

(14)

The calculation of the b values also differs from the mixing by parameters
(15).

b = 0.077796 RTc ( i , j ) pc ,( i , j )

The mixture critical properties follow the rules of Harstadt et. al.19 . Here Tc,(ii) = Tci and Tc (ij ) = TciTcj (1 kij ) for

i j with kij as binary interaction parameter, which is set to kij = 0.07447 after Tsonopoulos and Heidman 20 for the
mixing of hydrogen and oxygen. The critical pressures are evaluated as pc,(ii) = pci and pc (ij ) = Z c ( ij ) RTc ( ij ) Vc ( ij ) for

i j , with Z c (ij ) = 12 ( Z ci + Z cj ) , Vc (ij ) = 18 Vci13 + Vcj13 and c (ij ) = 12 (ci + cj ) for i j


3

Kinematic diffusivity
Binary diffusion coefficients for dilute gases (Fuller et. al.6)
In CFX the kinematic diffusivity is calculated by default from the bulk viscosity. As the diffusivity of hydrogen
is much higher than that of oxygen and water vapor the assumption of constant diffusivities underestimates the
hydrogen diffusion greatly.
Fuller et. al. modified the Chapman Enskog theory7 for the calculation of binary diffusion coefficients and
developed the empirical correlation
DAB =

0.0143T 1.75

(16)

pM AB2 ( v ) A + ( v ) B

with M AB = 2 /[(1/ M A ) + (1/ M B )] , where MA and MB are the molar weights of component A and B in [g/mol]
and v is the atomic diffusion volume which is empirically determined (O2: 16.3, H2: 6.12, H2O: 13.2). DAB is the
binary diffusion coefficient in [cm2/s]
Mixture diffusion coefficients for dilute gases
The mixture diffusion coefficient of one component into the mixture can be calculated using Blancs law8:
1

Di ,m = y j Dij
j =1, j i

5
American Institute of Aeronautics and Astronautics
092407

(17)

Diffusion coefficients for dense gases (Takahashi 9)


The low pressure diffusion coefficients are extended to high pressures using the correlation of Takahashi.

( Dp ) R = ( Dp ) R ,l (1 ATr B )(1 CTr E )

(18)

Here (Dp)R is the reduced value of the product of diffusion coefficient D


and pressure p to low pressure values ( (Dp)R = (Dp)/(Dp)0 ). A, B, C, E and
(Dp)R,l are empirical coefficients. Mixture critical values Tc and pc are
calculated as described below:
Tc' = y1Tc1 + y2Tc 2
pc' = y1 pc1 + y2 pc 2

The simple averaging by the molar fraction near critical pressures is not a
very good approximation.
Figure 1. Takahashi correlation 9 for
the reduced product of diffusion factor
and pressure to low pressure values
Molecular (dynamic) viscosity
Rigid Non Interacting Sphere Model
The default treatment of the viscosity of a gas in CFX is to assume that it consists of nonattracting rigid spheres
of diameter . Molecules move in the gas and transfer momentum or energy in a velocity or temperature gradient
exclusively by collisions. The viscosity for those rigid, non-interacting spheres is given by Hirschfelder et. al.10 as

= ( 26.69 10

( MT )

(19)

where is the molecular viscosity in [P = Poise = g/cm s ], M is the molecular weight in [g/mol], T is the
temperature [K] and s is the hard-sphere diameter in [angstrom].
Method of Chung et. al.11
Dilute Gas Viscosity
The theory of Chapman and Enskog7 accounts additionally for intermolecular forces by introducing an integral
expression * for a potential energy function describing interactions between molecules. The dilute gas viscosity is
described by

0 = ( 26.69 10

( MT )

(20)

Chung et. al. used the Chapman-Enskog theory 7 for the description of the dilute gas viscosity. The reduced
collision integral * depending on the intermolecular potential was estimated by an empirical equation developed
by Neufelder et. al. and the potential distance parameter was estimated by = 0.809 Vc with the critical molar
1

volume Vc in [cm /mol].

6
American Institute of Aeronautics and Astronautics
092407

As the Chapman-Enskog theory applies only for monoatomic gases, it was extended by a correction factor Fc
(Eq. 21) making it applicable for polyatomic molecular gases like hydrogen and oxygen. The correction factor
depends on Pitzers acentric factor , the dimensionless dipole moment r (Eq. 22) and for polar gases the a
correction factor for the hydrogen-bonding .
Fc = 1 0.2756 + 0.059035 r4 +

(21)

r = 131.3 (VcTc )

(22)

The dynamic viscosity for dilute gases can therefore be described as

0 = ( 4.0785 10

( MTc )

)V

Fc

(23)

Dense Fluid Viscosity


For dense fluids Chung et. al. extended the their theory for dilute gases to account for the effects of temperature
and pressure. The molecular viscosity for dense fluids is described by the empirical equation 24.

= + p

(24)

where = 0 [(G2 ) 1 + E6 y ] account for the temperature effects and p describes the pressure effects und the
molecular viscosity p = (36.344 106 )(( MTc ) Vc ) E7 y 2G2 exp( E8 + E9 (T *) 1 + E10 (T *)2 ) . The coefficients
1

Ei = ai + bi + ci r4 + di as well as the correction factor Fc are functions of the acentric factor, the reduced dipole

moment and a hydrogen bonding correction. The coefficients ai to di are empirically determined. For G1, G2 , y and
T* applies G1 = (1 0.5 y ) /(1 y )3 , G2 = [ E1 ((1 exp( E4 y )) / y ) + E2G1 exp( E5 y ) + E3G1 ] /[ E1 E4 + E2 + E3 ] , y = Vc 6
and T * = 1.2593Tr .
Thermal conductivity
Modified Eucken Model 12
Eucken proposed to calculate the contribution of translational and internal degrees of freedom separately for
polyatomic gases
103 M
1.77 R
= 1.32 +
cv
cv R

(25)

with the dynamic viscosity in [Ns/m2], M the molar weight in [kg/mol], the thermal conductivity in
[W/m K], specific heat capacity at constant volume cv in [J/mol K] and R the universal gas constant [J/molK]
Method of Chung et. al.11
Dilute gas thermal conductivity
The thermal conductivity for dilute gases is described by Chung et. al. as
0

0 = 7.452

7
American Institute of Aeronautics and Astronautics
092407

(26)

Where

is

= 1 + [(0.215 + 0.28288 1.061 + 0.26665Z ) /(0.6366 + Z + 1.061 )]

= 0.7862 0.7109 + 1.3168 and Z = 2 + 10.5 Tr2 , 0


conductivity for dilute gases in [cal/(cm s K)].
2

= cv R ,
is the dilute gas viscosity in [P] and 0 is the thermal
with

Dense fluid thermal conductivity


For dense fluids Chung et. al. extended the their theory for the thermal conductivity similar to the dynamic
viscosity. The thermal conductivity for dense fluids is described by the empirical equation 27.

= + p

(27)

where = 0 [( H 2 ) 1 + B6 y ] accounts for the temperature effects and p describes the pressure effects und the
thermal conductivity p = (3.039 104 )(( MTc ) Vc ) B7 y 2 H 2Tr . The coefficients Bi are calculated in the same way
1

as the coefficients Ei in the dynamic viscosity ( Bi = ei + fi + gi r4 + hi ) where ei to hi are determined empirically. y


and H2 are calculated from, y = Vc 6 and H 2 = [ B1 ((1 exp( B4 y )) / y ) + B2G1 exp( B5 y ) + B3G1 ] /[ B1B4 + B2 + B3 ] .

III. Experimental validation configuration


In the experiment, used for the validation of the code, the test case RCM-3 presented on the 2nd IWRCM 2. In
this test case, the single element combustor Mascotte V03 is fed with
liquid oxygen at 85 K and gaseous hydrogen at 287 K. This version
of the test rig, in use since 1997, allows operating conditions at
pressures above the critical pressure of oxygen (50.4 bar). Pressures
up to 100 bar can be achieved in the combustion chamber1,
representing supercritical conditions for hydrogen and transcritical
conditions for oxygen (T<Tcritical). The operating conditions used for
the workshop test case are summarized in Table 1.
The injector (Fig. 2) consists of an oxygen injector element with a
diameter of 3.6 mm at the inlet, diverging (with an angle of 8
between rotation axis and injector contour) to a diameter of 5 mm at
the exit. Hydrogen is injected coaxially through an annulus with
Figure 3. Mascotte high pressure
inner diameter of 5.6 mm
combustor V031
and outer diameter of
10 mm 3. The high pressure combustion chamber shown in figure 3 is a
square duct of 50 mm inner dimension and a length of 458 mm. At its
downstream end is a nozzle of variable shape, having a convergent length
of 20 mm and a throat
diameter of 9 mm
allowing
a
chamber
pressure of about 60bar2.
The chamber is made of
Figure 2. Mascotte single coaxial
stainless steel and fitted
3
injector element
with 4 silica windows for
optical access. Their internal side is cooled by a gaseous helium
film. The two lateral windows are 50 mm high and 100 mm long for
visualization. The upper and lower ones are also 100 mm long, but
only 10 mm wide. They allow a longitudinal laser sheet entrance
and exit1.
For the investigations the chamber was modeled rotationally
Figure 4. 20-sector of a rotational
symmetric with a radius of 28.81 mm in order to reproduce the
combustion chamber used for first CFD
internal chamber volume. At the end of the chamber a nozzle was
simulations.
fitted with a minimum radius of 15 mm in order to avoid backflow
8
American Institute of Aeronautics and Astronautics
092407

at the end of the domain. The injector was modeled realistically and was given a length of 50 mm in order to receive
a fully turbulent flow profile. In the simulation of the RCM-3 test
Conditions
H2
O2
case the film cooling of the windows is neglected. A sector of
Pressure
[MPa]
6
6
circumferential extent of 2 was modeled to receive a quasi twoMass flow rate [g/s]
70
100
dimensional model meshed with one hexahedral element layer
Temperature [K]
287
85
which is the way to set up two-dimensional problems in ANSYSDensity [kg/m3]
5.51
1177.8
CFX, since no genuinely 2-D solver exists in this code. The
Velocity [m/s]
236
4.35
domain was resolved by 200 x 1700 elements (radial x axial
direction). As inlet boundary conditions the mass flow rates
Table 1. Operating conditions for the
provided by the test case (Table 1) were set. At the outlet a
Mascotte test case RCM-3. 2
pressure of 60 bar was prescribed. The walls of the chamber were
assumed to be smooth and adiabatic.

IV. Results and Discussion


A. Combustion modeling
At the start of the investigations of this test case the combustion
was modeled using a simple eddy dissipation model with one-step
chemistry (2H2+O2 2H2O). This model results in temperatures
values up to 4200K, which is much higher than the adiabatic flame
temperature for the combustion of hydrogen and oxygen of 3600 K.
The temperature distribution in the flame can be improved by
the application of a flamelet model taking more species into
account. Here the representative interactive flamelet model
(Peters 17/18), implemented in CFX (CFX-RIF) was applied. This
Figure 5. Comparison of temperature
model considers in addition to the main species the radicals H, O,
distribution resulting from different
OH, H2O2 and HO2. Using CFX-RIF it is not possible to prescribe
combustion models
pure oxygen as
oxidizer,
there
must always be a mixture of N2 and O2. The mass fraction of N2 was
set to the minimum value which can be chosen, which is 0.001. The
maximum temperatures resulting from the flamelet model are about
1000 K lower than from the eddy dissipation model. With a value of
3200 K the maximum temperature is much more realistic. As it can
be seen in figure 5 the choice of the combustion model has only
small effects on the flame shape.
In figure 6 chemiluminescence measurements of the OH radical
Figure 6. Comparison of resulting OHare compared to the OH distribution of the simulation. The flame in
distribution with chemiluminescense
the CFD simulation
measurements 13
is longer than in
the
experiment.
Also, its spreading is not predicted correctly by the CFD, right
downstream the injector the spreading angle is to small, and the
maximum radial expansion is a little too large. A quantitative
comparison of the OH distribution is not possible, as is cant be
measured by the measuring method applied. Even qualitatively the
regions of maximum OH concentration in simulation do not fit the
experiment well, where the maximum values can be found in the
very thin combustion region right downstream the injector. This
mismatch could possibly be explained by the neglect of the species
OH* resulting from molecular collisions.
Comparing the axial temperature distribution of the flames
resulting from RIF and EDM model, it can be recognized that the
position of maximum temperature indicating the flame tip is in
Figure 7. Comparison of axial
about the same axial place for both combustion models, but it is
temperature distribution of experiment
9
American Institute of Aeronautics and Astronautics
092407

more downstream of the nozzle. The eddy dissipation model overestimates the temperature in the flame tip, whereas
it is significantly underpredicted by the flamelet model (Fig. 7).
As the material properties models in CFX are very simple, there is often a significant mismatch with the real
material properties, especially in the trans- and supercritical region. This might influence the flame length and
temperature. The effects of these simplifications are investigated below.
B. Diffusion effects
As hydrogen is a much more diffusive substance than water vapor
and oxygen, the mixing process as estimated using the bulk viscosity
for calculating the kinematic diffusivity (CFX default) might be
misrepresented here. In the investigations described below the simple
one-step eddy dissipation model had to be used, because the definition
of a diffusion coefficient for single components is not possible in
combination with the flamelet model in CFX. For hydrogen and
oxygen the diffusion coefficient into the mixture of both other
components was modeled using the method of Fuller et. al. in
combination with the Takahashi correlation (Eq. 16 and 18). It was
hoped that the increased
mixing would improve
the flame spreading
angle right downstream
of the injector and it
would shorten the flame
Figure 8. Comparison of axial oxygen
due to faster mixing and
distribution of different diffusivity
combustion.
Figure 8 shows the axial oxygen distribution. It can be seen that
using the bulk viscosity, the region pure oxygen persists slightly
farther downstream than using the new model.
Figure 9. Comparison flame resulting
Unfortunately Figure 9 shows that the effect of the higher
using different formulations for the
diffusivity is negligible compared to the turbulent diffusivity and the
kinematic diffusivity.
flame shape stays almost identical.
C. Modeling of Educts
Also for hydrogen, there are deviations of ideal gas behavior at moderate temperatures due to the high pressures.
Therefore hydrogen and oxygen were modeled using a real gas equation of state. All combustion products are
modeled as ideal gases due to the high flame temperatures.
Density and Heat capacity

Figure 10. Oxygen density at


constant pressure of 60 bar form
Ideal, Redlich Kwong and Peng
Robinson equation of state compared
with NIST 15

Figure 11. Hydrogen density at


constant pressure of 60 bar form
Ideal, Redlich Kwong and Peng
Robinson equation of state compared
with NIST 15
10
American Institute of Aeronautics and Astronautics
092407

For the calculation of the


density of hydrogen and
oxygen there are two different
real gas equations of state
available in CFX, the
Redlich - Kwong and the
Peng - Robinson
equation.
The densities resulting from
these
descriptions
are
compared to the density given
by the NIST15 chemical data
base and the one using the
ideal gas equation of state.
For oxygen the Redlich
Kwong equation fits the
NIST15 data best. The Peng

Robinson equation over-estimates the density by about 200 kg/m3 at the injection temperature of oxygen (T = 80K),
a deviation of 15 % (Fig. 10). This leads to an underestimated velocity at the orifice and therefore to a too small
momentum flux resulting in a shorter flame. The ideal gas formulation totally mispredicts the O2 density near and
below the critical temperature. As Figure 13 shows, the density of hydrogen is very well estimated by all equations
of state.
Another material property
which might affect flame
shape and temperature is the
heat capacity. Figure 12
shows that for oxygen the heat
capacity resulting from the
Redlich Kwong approach
overestimates the NIST data
by a factor of 2 below the
critical temperature. The cold
oxygen jet is heated up slower
leading to an overestimated
density which might affect the
Figure 13. Hydrogen specific heat
Figure 12. Oxygen specific heat
flame length in the opposite
capacity at constant pressure of 60
capacity at constant pressure of 60
way as the density. The heat
bar form Redlich Kwong and Peng
bar form Redlich Kwong and Peng
capacity of hydrogen is rather
Robinson
equation
of
state
compared
Robinson equation of state compared
15
15
well estimated for both
with
NIST
data.
with NIST data.
equations of state.
Figure 14 shows, that the overestimated density resulting
form the Peng-Ronbinson formulation is the most important
influence to the flame length of the model improvements
tested. The flame tip is a little further upstream, but the axial
position of the maximum radial extension stays almost the
same. Unfortunately the slight improvement of the match with
the experiment by using the Peng-Robinson equation occurs for
the wrong reason, because the shortening of the flame is
reached by an oxygen density which is not physically valid.
Figure 14. OH distribution in the flame form
Molecular (Dynamic) Viscosity and Thermal Conductivity
Redlich Kwong and Peng Robinson equation.
The dynamic viscosity affects the stress tensor describing internal friction within the fluid. The viscosity behaves
very differently in low and high-pressure regions. The transcritical region is very special; here the viscosity at
constant pressure no longer
decreases with temperature,
as it does in a gas. It
behaves like in a liquid and
rises rapidly with falling
temperature. In CFX very
simple
models
are
implemented
for
the
description of this transport
property, which are not
valid at very high pressures.
A rigid non-interacting
sphere model is used,
Figure 16. Hydrogen dynamic
Figure
15.
Oxygen
dynamic
which applies only for
viscosity form RISM and Chung et.
viscosity form RISM and Chung et.
mono-atomic gases at very
al. compared to NIST data. 15
al. compared to NIST data. 15
low
pressures.
It
is
compared to the empirical model of Chung et. al.11 which was implemented into CFX in order to account for the
effects of the simplifications in this material property. Figures 15 and 16 show the viscosities resulting from both
models compared to NIST15 data.
11
American Institute of Aeronautics and Astronautics
092407

The oxygen data are fit very well by the Chung correlation. In contrast, the RIS model is not able to predict the
rise in the viscosity below the critical temperature. The experimental
hydrogen
data
are
represented
qualitatively
rather well, the exact value
is not met by the Chung
model, but the fit is much
better than with the simple
model.
Also for the thermal
conductivity a very simple
modified Eucken model is
implemented in CFX. This
model is compared to an
Figure
17.
Oxygen
thermal
empirical
model
also
Figure 18. Hydrogen thermal
conductivity
form mod. Eucken
developed
by
Chung
et.
al.
conductivity form mod. Eucken
model and Chung et. al. compared to
Figures
17
and
18
show
the
15
Model and Chung et. al. compared to
NIST data.
predicted
values
for
NIST data. 15
hydrogen and oxygen. The
O2 thermal conductivity is reproduced rather well by the Eucken
model, but in very cold regions the values are highly overestimated.
The agreement for hydrogen is again much better for the Chung
method than for the Eucken model, but there are still significant
deviations.
Figure 19 compares the OH-distributions within the flames
resulting from the application of the simple models and that
calculated using the improved empirical correlations. There is
almost no difference in shape.

Figure 19. Comparison of flame shape


resulting from default models compared to
improved models

D. Turbulence modeling
Since it was not possible to improve the flame shape using improved representations of material data it was tried
to improve the agreement by a modification of the turbulence model, since the spread of round jets is known to be
mispredicted by almost all two-equation turbulence models.
Effect of turbulent Prandtl number
The turbulent Prandtl number is set by default to a value of 0.9 in CFX.
It is coupled to the turbulent Schmidt number in order to treat turbulent
diffusion processes for heat and species identically.
It is known from the modeling of aircraft turbine engines that
Prt 0.3 - 0.5 sometimes improve the prediction of the temperature field.
Therefore Prt was varied in order to investigate the effect on the resulting
flame.
Figure 20 shows that a reduction of the turbulent Prandtl number from
0.9 to 0.7 leads to a slight shortening of the flame and to a reduction of the
temperature. However this effect is much too small to match the
experimental results. If the turbulent Prandtl number is reduced further
(Prt = 0.5), the flame evolutes an even more unrealistic shape.

Figure 20. Effect of the variation of


the turbulent Prandtl number

12
American Institute of Aeronautics and Astronautics
092407

Effect of modification of model constants C1 and C2


All simulations above were performed using a simple standard k--model (C1 = 1.44, C2= 1.92). It is known
from literature that this model overestimates the turbulent kinetic
energy in the simulation of axisymmetric free jets. Gaillard et. al. 16
showed that a choice of C1 = 1.52 and C2= 1.89, yielded a reasonable
fit to their experimental results. Applied to the current case, this
method led to even
worse results. Figure 21
compares
flames
resulting
from
the
application
of
the
standard
k--model
Figure 21. Temperature distribution
compared
to
the
using EDM resulting from Gaillard (1),
Gaillard approach and a
standard (2) and modified (C1=1.3) k-simple modification of
model (3)
the C1 coefficient to a

Figure 23. Comparison of axial


temperature distribution with calculated
distributions resulting from standard and
modified k--model for the two different
combustion models

value of 1.3. The third


approach seems to
improve
the
flame
shape.
Figure 22 shows the
Figure
22.
Comparison
of
comparison of CFD and
chemiluminescence measurements of the
experiment for the
OH
distribution
with
calculated
standard
and
the
distribution resulting from standard and
modified model with a
modified k--model
C1 value of 1.3. The
OH distribution from the modified model fits the experimental results
almost exactly. Figure 23 compares the experimental temperature
distribution with those resulting from simulations using different
combustion and turbulence models. The position of the flame tip is
met rather well using the modified turbulence model (C1 = 1.3). The
temperature almost agrees with the simple one-step eddy dissipation
model and is highly underestimated by the flamelet model which
actually is the more physically valid model.

E. Evaluation of mixing rules


The most serious simplification in CFX is the assumption of
ideal mixing, even in high pressure and low temperature real gas
applications. This cannot be changed currently, since there is no
direct user access to the internal mixing procedure. This assumption
leads to a significant misrepresentation of the mixture density [Fig.
23]. Here a mixture of a hydrogen and oxygen is investigated with
an O2 molar fraction of 75 %. The mixture density provided by CFX
is compared to common mixing rules by the critical properties, or
by the parameters in the case of the Peng - Robinson equation.
These two approaches provide very similar results which differ
remarkably from the mixing by molar volumes and even more from
the mixing by partial volumes which is implemented in CFX. This
enormous underestimation leads to an overestimation in the
momentum flux and is therefore probably the cause for the
mismatch in the flame length.

Figure 23. Comparison of mixture density


resulting from different real gas mixing rules
with CFX results

13
American Institute of Aeronautics and Astronautics
092407

V. Conclusion
In this paper, transcritical real gas combustion is simulated using the commercial CFD solver ANSYS CFX with
improved material models and validated using the Mascotte RCM-3 test case
The simple one-step eddy dissipation model yields temperatures, of up to 4200 K, which are much higher than
the adiabatic flame temperature of hydrogen and oxygen (~ 3600 K). This has been improved using the flamelet
model (CFX-RIF) implemented in CFX. This model provides reasonable maximum temperatures of 3200 K which
are, however, even lower than the maximum temperature measured by Bouchardy et. al.14.
In CFX diffusion coefficients are calculated using the bulk viscosity. This approach assumes that every
substance diffuses into the other substances with the same rate. As hydrogen is a very diffusive substance, diffusion
was modeled using a special empirical approach in order to investigate the effects on the flame. Although the effects
of the different modeling can be seen in the oxygen distribution, the flame is almost not affected.
The propellants were described using the Redlich-Kwong equation of state. This approach provides a very good
agreement in the density but it mispredicts the oxygen heat capacity below the critical temperature by a factor of 2.
Using the Peng-Robinson equation instead, the heat capacity fits the NIST data very well but the density of oxygen
is overpredicted by about 15 %. The resulting slight shortening of the flame is due to this significant deviation from
the true material data, so it cannot be accepted as valid model improvement.
Transport properties are modeled very simply, in CFX, leading to mismatches especially in the trans- and
supercritical region. This was improved for oxygen and hydrogen using an empirical approach of Chung et. al.
Viscosity and thermal conductivity of both materials could be reproduced very well, but the effects on the flame
were negligible.
Finally it was tried to fit the CFD to the experiment modifying the turbulence model. First the turbulent Prandtl
number was varied as it is known from the field of aircraft turbines that this model constant is very dependent on the
application. By a sensible modification of this constant no satisfying fit to the experimental results could be found.
From the literature (Gaillard) it is known that for axisymmetric free jets the k--turbulence model overestimates
the turbulent kinetic energy. Gaillard et. al. found that the C1- coefficient has to be set to 1.52 and the C2coefficient to 1.89 to reach a proper agreement between simulation and experiment. Unfortunately this approach
leads to an even worse result here. A modification of the C1- coefficient to 1.3 provided a very good fit with the
experimental results. No physical explanation can be given, but this approach seems to fit the experiment well using
CFX.
The main reason why it is not possible to reach a better agreement is probably the assumption of an ideal mixing
process for the different components. As figure 24 shows there is an enormous misprediction of the mixture density,
which doubtless has significant effects on the flame. The assumption of ideal mixing is very difficult to work around
in CFX. LES simulations using an in-house-code with proper material models might improve the simulation.

Acknowledgments
The validation of a CFD solver always needs experimental results. Lucien Vingert supported this work by
providing those results of the Mascotte test facility in a very uncomplicated way, which we are very grateful for.
Also the cooperative support of the ANSYS team providing valuable information about CFX solver and
modelling details is gratefully acknowledged.

References
1
Habiballah, M., Orain, M., Grisch, F., Vingert, L., and Gicquel, P., Experimental studies of high-pressure cryogenic flames
on the Mascotte facility Combustion Science and Technology, Vol. 178, 2006, 101-128.
2
Habiballah, M., Zurbach, S., Test Case RCM-3, Mascotte single injector Proceedings of the 2nd International Workshop
on Rocket Combustion Modeling, Lampoldshausen 2001
3
Vingert, L., Nicole, A., Habiballah, M., Test Case RCM-2, Mascotte single injector Proceedings of the 3rd International
Workshop on Rocket Combustion Modeling, Vernon, France 2006
4
Vingert, L., Habiballah, M., Vuillermoz, P. and Zurbach, S., Mascotte, a test facility for cryogenic combustion research at
high pressure 51st International Astronautical Congress (IAF), 2000.
5
CFX, Software Package, Ver. 11.0, ANSYS Germany, 2007.
6
Fuller, L. N., Schettler, P. D. and Giddings, J. C., A new method for prediction of binary gas-phase diffusion coefficients,
Industrial and Engineering Chemistry, Vol. 58, No. 5, May 1966.
7
Chapman, S. and Cowling, T. G., The mathematical theory of nonuniform gases, 2nd ed., p.245, Cambridge University
Press, London, 1952.
8
Blanc, M. A., J. Phys. (Paris), Vol. 7, 1908, 825.

14
American Institute of Aeronautics and Astronautics
092407

9
Takahashi, S., Preparation of a generalized chart for the diffusion coefficients of gases at high pressures, Journal of
Chemical Engineering of Japan, Vol. 7, No. 6, 1974.
10
Hirschfelder, J. O., Curtiss, C. F. and Bird R. B. Molecular theory of gases and liquids, Wiley, New York, 1954.
11
Chung, T.-H., Ajlan, M., Lee, L. L. and Starling, K. E., Generalized Multiparameter Correlation for Nonpolar and Polar
Fluid Transport Properties, Industrial & Engineering Chemistry Research , Vol. 27, No. 4, 1988.
12
Svehla, R. A., Estimated Viscosities and Thermal Conductivities of Gases at High Temperatures, NASA Tech. Rept., R132, Lewis-Research Center, Cleveland, Ohio, 1962.
13
Candel, S., Herding, G., Synder, R., Scouflaire, P., Rolon, C., Vingert, L., Habiballah, M., Grisch, F., Palat, M.,
Bouchardy, P., Stepowski, D., Cessou, A., Colin, P., Experimental Investigation of Shear Coaxial Cryogenic Jet Flames,
Journal of Propulsion and Power, Vol. 14, No. 5, 1998, 826-834.
14
Bouchardy, P., Grisch, F., Clauss, W., InternationalWorkshop on Research Status and Perspectives in Liquid Rocket
Combustion Chamber Flow Dynamics, Paris, 1999, p.345.
15
E.W. Lemmon, M.O. McLinden, D.G. Friend. Thermophysical Properties of Fluid Systems. NIST Chemistry WebBook,
NIST Standard Reference Database Number 69, National Institute of Standards and Technology, Gaithersburg MD, 20899
(http://webbook.nist.gov), 2005
16
Gaillard, P., Multidimensional Numerical Study of the Mixing of an Unsteady Gaseous Fuel Jet with Air in Free and in
Confined Situations, International Congress & Exposition, Detroit, Michigan, February 27 March 2, 1984.
17
Peters, N., Turbulent Combustion, Cambridge University Press, Cambridge, 2002.
18
Peters, N., Laminar Diffusion Flamelet Models in Non-Premixed Turbulent Combustion, Prog. Energy Combust. Sci.,
Vol. 10, 1984, 319-339
19
Harstad, G., Miller, R. S., and Bellan J., Efficient high pressure equations of state, AIChE Journal, Vol. 43, No. 6, 1997,
703-723.
20
Tsonopoulos, C. and Heidman J. L., High pressure vapor-liquid equilibria with cubic equations of state, Fluid Phase
Equilibria, Vol. 29, 1986, 391-414.
21
Miller, R. S., Harstad, G. and Bellan, J., Direct numerical simulations of supercritical fluid mixing layers applied to
heptane-nitorgen, J. Fluid Mech., Vol. 436, No. 6, 2001, 1-39.
22
Oefelein, J., Mixing and combustion of cryogenic oxygen-hydrogen shear-coaxial jet flames at supercritical pressure,
Combustion Science and Technology, Vol. 178, 2006, 229-252.
23
Miller, R. S., Harstad, G. and Bellan, J., Evaluation of equilibrium and non-equilibrium evaporation models for manydroplet gas-liquid flow simulations, International Journal of Multiphase Flow, Vol. 24, 1998, 1025-1055.
24
Yang, V., Modeling of Supercritical Vaporization, Mixing, and Combustion Processes in Liquid-Fueled Propulsion
Systems, Proceedings of the Combustion Institute, Vol. 28, 2000, 925-942.
25
Delplanque, J.-P. and Sirignano, W. A., Numerical study of the transient vaporization of an oxygen droplet at sub- and
super-critical conditions, International Journal of Heat and Mass Transfer, Vol. 36, No. 2, 1993, 303-314.
26
Daou, J., Haldenwang, P. and Nicoli, C., Supercritical Burning of Liquid Oxygen (LOX) Droplet with Detailed
Chemistry, Combustion and Flame, Vol. 101, 1995, 153-169.
27
Zong, N. and Yang, V., Cryogenic Fluid Jets and Mixing Layers in Transcritical and Supercritical Envirionments,
Combustion Science and Technology, Vol. 178, 2006, 193-227.
28
Cheng, G. and Farmer, R., Real Fluid Modeling of Multiphase Flows in Liquid Rocket Engine Combustors, Journal of
Propulsion and Powers, Vol. 22, No. 6, Nov.-Dec. 2006, 1373-1381.
29
Benarous, A., Liazid, A. and Karmed, D., H2/O2 Combusition under Supercritical Conditions, Third European
Combustion Meeting ECM, 2007.

15
American Institute of Aeronautics and Astronautics
092407

Potrebbero piacerti anche