Sie sind auf Seite 1von 19

Improvements in Analyzing High-Speed Fuel/Air Mixing

Problems Using Scalar Fluctuation Modeling


S. Mattick
*

Combustion Research & Flow Technology, Inc. (CRAFT Tech)
Huntsville, AL
Phone: 256-883-1905/Fax: 256-833-1967
Email: stephen@craft-tech.com

and
K.W. Brinckman

, S.M. Dash

, and Z. Liu

,
Combustion Research & Flow Technology, Inc. (CRAFT Tech)
Pipersville, PA 18947
Phone: 215-766-1520/Fax: 215-766-1524
A scalar fluctuation model (SFM) that solves partial differential equations for
energy/species variance and corresponding dissipation rates is presented, along with several
applications to high-speed fuel/air mixing problems. The model is implemented in a k-
epsilon turbulence model framework with unified compressibility and low Re extensions,
specialized for high speed aero-propulsive flow applications. It is used to predict the spatial
variation of turbulent Prandtl and Schmidt numbers using time-scale relations, providing
more accurate and reliable solutions than those based on user-specified average-values. Over
the past several years, the authors and coworkers have systematically upgraded the SFM to
treat flows of increasing complexity, using a building-block approach to ensure that
modifications made to improve the analysis of more complex cases will not degrade the
model performance in analyzing fundamental cases. A GUI-driven building-block data base
(BBDB) tool has been developed to facilitate the validation/calibration process, which
contains the various data sets we are working with (experimental and LES) along with grids
and solution files, and scripts to take CFD output and put it into the format required to
compare with the data. This paper will describe the latest version of the SFM, its
application to select fundamental cases in the BBDB, and a detailed description of its
analysis of the SCHOLAR fuel/air mixing/combustion data in which we have examined grid
resolution sensitivities and compared results using the SFM with those using different values
of constant turbulent Prandtl and Schmidt numbers.
I. INTRODUCTION
Accurate modeling of scalar transport, such as internal energy and species concentration, is needed to properly
predict the temperature and fuel/air distribution in high-speed aero-propulsive devices, and thus to predict
combustion efficiency and overall performance. The effect of turbulence plays a dominant role in scalar transport
and is accounted for in a RANS framework through the application of a turbulent Prandtl (Pr
t
) and turbulent
Schmidt (Sc
t
) number for thermal and species mixing, respectively. It is now well recognized that the application of
constant-average values of these numbers for complex flows is inadequate since: different values apply to different
very basic flows (e.g., 0.7 is used for round jets, 0.9 is used for boundary layers); these numbers vary widely in
complex flows; and, the values differ with varying levels of compressibility in the flow. Turbulence predictions for
high speed flows are complicated by compressibility effects which have a first-order influence on fundamental

*
Research Scientist, 3313 Memorial Parkway South, Suite 108, Member AIAA.

Senior Research Scientist, 6210 Kellers Church Rd., Member AIAA

President & Chief Scientist, 6210 Kellers Church Rd., Associate Fellow AIAA.

Research Scientist, 6210 Kellers Church Rd., Member AIAA



American Institute of Aeronautics and Astronautics

1

mixing for basic (cold air/air) flows, and have an even more complex influence on flows with large density
variations produced by differences in temperature and species molecular weights.
To address such issues, a scalar fluctuation model (SFM) has been developed and systematically extended to
analyze high-speed aero-propulsive flow problems such as those encountered in scramjets, as well as in a variety of
related mixing problems such as those associated with missile plumes and divert/control jets. Brinckman et al.
1

extended earlier methodology formulated for lower speed flows for deriving local values of turbulent Prandtl
number in hot jets. Transport equations for both temperature and species mass-fraction variance were formulated
which incorporated both high-speed compressibility corrections and low Reynolds-number near-wall extensions.
Various simulations were performed for basic fuel/air mixing problems and for problems with more complex
geometries. Brinckman et al.
2,3
and Calhoon et al.
4
presented extended methodology which is capable of modeling
high-speed reacting flows. Validation cases were presented in Ref. [2],[3] and [4] which demonstrated good
agreement for both reacting and non-reacting flows. Brinckman et al.
3
showed model comparisons for a non-
reacting high-speed fuel injector, and demonstrated the complex three-dimensional variation in turbulent scalar
mixing characteristics associated with these types of configurations. Ott et al.
5
showed additional predictions for a
non-reacting angled fuel injector, including comparisons to LES predictions. With a scalar variance modeling
capability for variable Pr
t
and Sc
t
in place and validated against several fundamental data sets, the current work
seeks to: extend the approach used for validation using a new Building-Block Data Base (BBDB) tool; and, analyze
more complex 3D problems, such as the SCHOLAR combustor data, which typifies high-speed fuel injection and
combustion in a scramjet combustor type of environment.
II. SFM TURBULENCE MODEL EQUATIONS
A. k- Turbulence Modeling Framework
The k- turbulence model used for aero-propulsive flows in the CRAFT CFD

Navier-Stokes code is based on


the original model for free shear flows of Launder et al.
6
with enhancements incorporated by Papp and Dash
7
to
account for compressibility effects in a unified manner, following concepts introduced by Sarkar and Zeman
8,9

along with other specialized extensions (e.g. vortex stretching, etc). The near-wall damping formulation of So et al.
10

is used with modifications that permit our application of the original Launder et al.
6
coefficients and not those of So
et al.
10
Details, plus model equations are provided in Ref. [2]. The scalar fluctuation model (SFM) employed in the
CRAFT CFD

code is based on extensions to the temperature variance model of Sommer et al.


11
. It includes a two-
equation species mixture fraction variance k
f
-
f
model used for obtaining the turbulent Schmidt number, and a two-
equation energy variance k
e
-
e
model used for obtaining the turbulent Prandtl number, via time-scale relations.
While the added expense of solving transport equations for both species and energy variance can be avoided if a
constant value of turbulent Lewis number is assumed,

t
t
t
Pr
Le
Sc
= (1)
this assumption is not generally valid, and the Lewis number can vary significantly in many of the flows of interest.
Figure 1 shows how the Schmidt number can vary in an angled (transverse) fuel injection flowfield. Prandtl number
variations governed by temperature/internal energy gradients will be quite different and Lewis number variations in
most fuel injection flowfields are very substantive.

Figure 1. Schmidt Number variations in a Transverse Fuel Injection Flowfield with High-Speed Approach
Flow.


American Institute of Aeronautics and Astronautics

2

B. Energy Variance Equations
The energy variance k
e
-
e
formulation used to predict local values of turbulent Prandtl number has a production
source term for the energy variance k
e
=<ee>based on the gradient of the internal energy, e. Earlier versions had
used the gradient of temperature for the production term, but for chemically reacting flows, use of energy variance is
preferable
4
. The transport equations for energy variance and its dissipation rate
e
are:

2
,
( )
( )
2
e
e e
e
k e j
u k
k k e j
t

t
t x x x x
j j j

2



+ = + +


(2)

2
,
( )
( )
1 2

3 4 5
j e
e e
j j e j e
e e
e
e
u
e
t
C C
t d d
t x x x k k

C P C C
d k d d T
k k k

e
j
x


+ = + + +




+ + +


(3)
where the dissipation rate is defined as,

e
k k
e e
x x


=

(4)
and the heat transfer time scale is,

,
e
m e
e
k k


= (5)
The compressibility correction used in the k- model
7
has the relation,

2 2
P P M P M
k k 1 T k 2 T
=

(6)
In this equation, P
k
is the production term in the transport equation for k and
T
M

is the local turbulent Mach number


with respect to the sound speed, c, lagged by a value of =0.2, so that compressibility corrections do not initiate
until this fluctuation level is exceeded
7
. Thus,

2
2 / ,
T
M k c M max M , 0
T T

= =

(7)
In Eqn. (3),
T
is a near-wall damping function included to capture low-Reynolds number behavior. Details of
this term are provided in Ref. [1] for the temperature variance formulation. An analogous form is used for the energy
variance dissipation rate, with the scalar fluctuation time scale provided by the ratio k
e
/
e
. Constants used in the k
e
-
e

model, summarized in Table 1, are consistent with those proposed by Sommer et al.
11
with the exception of
, k

and
,e
, whose value of 1.0 has been found to provide better agreement with data. Wall thermal boundary
treatments using a wall-function approach
12
or a wall resolved approach
11
are both available.
The turbulent Prandtl number with near-wall damping is given by

e
T
e
C f
k
Pr
C f k

= (8)
and the turbulent thermal diffusivity is calculated from,

,
t
t
T
v
C f k
Pr
m e
= = (9)
where C

=0.14, which differs slightly from the value presented by Sommer et al.
11
to better simulate the turbulent
Prandtl number for jet mixing problems. The coefficients C

and f

are related to the k- turbulence model and are



American Institute of Aeronautics and Astronautics

3

discussed in Ref. [7]. For the nearwall damping function, f

, following Abe et al.


13
, we use the Kolmogorov
velocity scale u

= ()
1/4
to define a wall distance y* = u

y/
,
which can be applied in both attached and detached
flows. The expression for the near-wall damping function f

used in the current fluctuating-scalar model is



*
2
1
1/4
/
1
t
C
y A
f f e
T
Re


+

= +

(10)
Use of the internal energy as a basis for this model removes the chemical source terms from the transport
equations that appear for the temperature-based formulation. This model has also been formulated so that it reduces
exactly to the nonreacting formulation of Brinckman et al.
1
when the fluid mixture specific heats are constant. The
model predicts both low and high speed flows using a single set of calibration coefficients. For reacting flow
simulations using a pdf-based model, the required temperature fluctuations may then be extracted from k
e
as
described by Calhoon and Kenzakowski
14
.

Table 1. Scalar Variable Model Constants
C
d1
C
d2
C
d3
C
d4
C
d5
C
1
A
+

k,e

,e

2.0 0.0 0.72 2.2 0.8 0.1 45 1.0 1.0
C. Mixture Fraction Variance Equations
Transport equations for the mixture-fraction variance k
f
=<ff> and its dissipation rate
f
are similar to the
energy variance model, except that the source term on the k
f
equation is based on the mixture fraction f, where f is
generally taken to be the effective dilution ratio of an inert species (e.g., f=1 in the unmixed fuel jet and f=0 in the
surrounding airstream). The mixture fraction variance equations are given as:

2
,
( )
( )
2
f
f f
t
f
j j k f j j
u k
k k
D f j
D D
t
t x x x x

+ = + +


(11)

2
,
( )
( )
1 2

3 4 5
f
f f
j j f j f
f f
f
f
u D

f j
t
D D C C
t d d
t x x x k k


C P C C
d k d d f
k k k

f
j
x



+ = + + +





+ + +



(12)
where f is tracked as a conserved scalar. In non-reacting problems, the mass fraction of an individual species
behaves as a conserved scalar, and a mixture fraction can be derived by normalizing one of the species mass
fractions. This approach provides computational efficiency by eliminating the need to solve an additional transport
equation for mixture fraction. In reacting problems, the same approach can be used with any inert species, as its
mass must be conserved. In the absence of an inert species, an additional transport equation for the mixture fraction
itself must be solved to assure that the variance k
f
is based on a conserved scalar (see Ref. [2]).
Here, the dissipation rate is defined as

f
k k
f f
D
x x


=

(13)
and the mass transfer time scale, based on the turbulent velocity and species mixing time scales, is

,
f
m f
f
k
k


= (14)
The turbulent mass diffusivity is calculated from:

,
t
t
t
D C f k
Sc

m f
= = (15)

American Institute of Aeronautics and Astronautics

4

and the turbulent Schmidt number with near-wall damping is:

f
t
f
C f
k
Sc
C f k

= (16)
where the model parameters C

, C

, f

and f

are as described earlier for the energy variance model. The model
constants for the mixture fraction variance are those used in the energy variance model as given in Table 1.
III. Building-Block Data Base (BBDB) Tool
In calibrating and validating the SFM, a building-block approach is now being used where we start with very
basic cases and work our way forward analyzing cases of increasing complexity. If we need to add corrections or
modify the SFM for more complex cases to get good agreement with data, we have to go back to the more basic
cases and rerun them to ensure that the changes made do not degrade the earlier comparisons. This can be a very
time-consuming and tedious process, which we have now expedited by constructing a GUI-driven BBDB tool. This
tool has the basic features summarized in Table 2 below, with some of the cases we have included shown in Table 3.
Figures 3 and 4 show case descriptions, which summarize some of the cases included in the BBDB. For each case,
we have the data (experimental or LES) in line plot format and scripts for the CFD solution files that provide direct
graphical comparisons with the data with
no user intervention. Several cases can
be run sequentially, so that after one
makes some modifications to the SFM, a
sequence of runs can be performed
overnight with the comparisons available
in the morning, plotted as per the scripts
set up for each case. The BBDB tool has
also proven useful in comparing results
of different codes and we have been
developing code-specific interface
routines to facilitate making such
comparisons. Once we have completed
installation of the BBDB on a dedicated
web-site, outside users will be able to
use it as well, as shown schematically in
Figure 2.

Table 2. Building Block Data Base (BBDB) Tool

Tool kit / data base provides:
user-friendly / web-based environment for accessing and using validation case files
detailed description of validation case
files with available experimental or LES data
solution files ( using various turbulence models, etc)
grid file
restart file and instructions on how to restart (code specific)
current solution file: case specific processing (scripts)
- utilities to format data files for direct comparison
Data organized in a Building-Block manner
automated run sequencing is available so we can repeat a series of most relevant cases to see
how a modeling change impacts comparisons with data and earlier solutions


Figure 2. Schematic of Web-based BBDB System used for SFM
Validation.





American Institute of Aeronautics and Astronautics

5

Table 3. Select SFM Validation Data Sets and GUI Driven Data Retrieval



Figure 3. Beach Hydrogen/Air Coaxial Reacting Jet
Data Set.
Figure 4. SCHOLAR Angled Fuel Injection
Combustion Data Set.


American Institute of Aeronautics and Astronautics

6

IV. Fundamental Case Studies
A. Compressibility-Effects Air/Air Studies
The effects of compressibility on jet mixing problems are quite significant, and can be illustrated by the
comparisons shown below for a round, Mach 2 jet into still air (Seiner
15
), and for a transverse jet exhausting into a
supersonic approach flow with boundary layer (Spaid and Zukoski
16
). Including the unified compressibility
correction of Reference [7] into the k- turbulence model reduces the rate of jet decay, and increases the extent of the
transverse jet induced separation bubble, in agreement with the data, as shown in Figures 5 and 6 below.






Figure 5. Analysis of Seiner Mach 2 Jet into Still
Air with and w/o Compressibility-Correction (CC).
Figure 6. Analysis of Spaid and Zukoski Transverse Jet
Data with and w/o CC
B. Low Speed vs. High Speed Hot, Round Jet Behavior
Our studies using SFM have indicated fundamental differences in the behavior of subsonic (Figure 7a) and
supersonic (Figure 7b) hot jets. For hot subsonic jets, the velocity and temperature fluctuations behave in similar
manner as their intensities asymptote toward the same level after the mixing reaches the axis (as confirmed by
comparisons with data sets such as those of Lockwood and Moneib
17
). In contrast for supersonic jets, the velocity
fluctuations on the centerline asymptote, but the temperature fluctuations peak and then decay. The SFM predicted
trends are confirmed by comparisons with an LES calculation for the supersonic jet in Figure 7b.



(a). Hot Subsonic J et (b). Hot Supersonic J et
Figure 7. Low Speed vs. High Speed Hot, Round Jet Behavior, Fluctuation Intensity

American Institute of Aeronautics and Astronautics

7

C. Scalar Fluctuation Differences in Variable Density Subsonic and Supersonic Round Jets
As shown in the comparisons below (Figure 8)
performed using SFM, for subsonic jets with the same
density ratio (produced by either temperature or
molecular weight), the temperature and species
fluctuations are identical. This should be contrasted
with the behavior in supersonic jets where it is seen
that temperature fluctuation levels are significantly
higher than species fluctuation levels, and they peak
and decay rather than flattening after the mixing
reaches the axis. This difference in behavior
exemplifies how compressibility affects scalar
fluctuations, and modeling this behavior properly is a
matter of on-going research, as we look to analyze
more challenging supersonic flows, with shocks and
other complexities.

Figure 8. Subsonic and Supersonic Variable Density
Ratio Jets.
V. Basic Fuel / Air Mixing and Combustion Studies
A. Beach/Evans Coaxial H
2
/Air Jet Study
In the experiment of Evans et al.
18
, a Mach 2 hydrogen jet is exhausted into a co-axial stream of vitiated air at
Mach 1.9. This is a basic coaxial H
2
/air high-speed jet mixing and combustion case with no geometric complexities.
The jet issues from a nozzle with inner diameter d =0.6525 cm and outer diameter d
o
=0.9525 cm. A schematic of
the experimental apparatus and flow conditions for the inner and outer jets is provided in Figure 3. The case was
analyzed using a basic H/O reaction mechanism (7 species/8 reactions) with both the Prandtl and Schmidt numbers
prescribed (Pr
t
=Sc
t
=0.7), and using the SFM model. Figure 9(a) and (b) show the predicted variation of H
2
along
the jet centerline, and a profile of H
2
O at a location about 22 diameters downstream of the nozzle exit plane. The
predicted contours of Pr
t
and Sc
t
for this case are shown in Figure 10 and do not vary significantly, with values
ranging from about 0.55 - 0.75 in the developed jet. Using averaged values of 0.7 appears to do an adequate job for
this basic case.



(a). H
2
variation along jet Centerline (b). H
2
O profile at x/d ~22
Figure 9. Beach/Evans Coaxial H
2
/Air Jet, Comparison of Predictions to Data.



American Institute of Aeronautics and Astronautics

8



(a). Turbulent Prandtl Number (b). Turbulent Schmidt Number
Figure 10. Prandtl and Schmidt number variations for Beach/Evans Jet.

Figure 11. Schematic of experimental set-up
for the Burrows and Kurkov
19
combustor.
B. Burrows and Kurkov H
2
Slot Jet Ducted Mixing and Combustion Study
This experiment consisted of sonic hydrogen injection
through a backward facing step into a Mach 2.44 vitiated air
(or with N
2
in the non-combusting case) stream. Figure 11
presents a schematic of the experimental set-up. The flow
conditions for this case are summarized in Table 4. Contours
of Prandtl, Schmidt and Lewis number are shown in Figure
12. For the mixing case, the variations are seen to be modest,
and the analysis of this case with constant values has been
shown to be adequate. For the combusting case, note the
pronounced change in these transport parameters at the
ignition location which has led to markedly improved
comparisons with the data (see Refs. [2]-[4]). The response of
the SFM to the changes in turbulence produced by heat release due to combustion is an extremely promising feature
of this model and has led to improvements in comparisons with data for several case, including the SCHOLAR data
to be discussed below.

Mixing Experiment Combustion Experiment



Figure 12. Burrows & Kurkov Study Contours of Pr
t
, Sc
t
and Le
t
.


American Institute of Aeronautics and Astronautics

9

Table 4. Vitiated air and fuel exit conditions for Burrows and Kurkov
19
experiment.
Vitiated Air Hydrogen
Static Pressure 101.325 kPa 101.325 kPa
Static Temperature 1270 K 242 K
Mach Number 2.44 1.0
H
2
mass fraction 0.0 1.0
O
2
mass fraction 0.258 0.0
N
2
mass fraction 0.486 0.0
H
2
O mass fraction 0.256 0.0
VI. Analysis of SCHOLAR Data
A. Overview
The SCHOLAR data set (see Figure 4) is intended to validate CFD models in an environment which typifies that
in a scramjet combustor, with flush/angled injectors. The 3D SCHOLAR angled fuel injector mixing and
combustion experiments were conducted at NASA Langley Research Center Ref. [20-21], providing data for CFD
model validation, including mean temperature, wall pressure, and species measurements. These data sets are useful
for analyzing the capabilities of the SFM to capture turbulent mixing effects in more complex flows involving
hydrogen injection into three-dimensional cross-flows. Data was obtained at 4 stream-wise planes downstream of
the injector, with the X-location of each stream-wise plane provided in Table 5.

Table 5. X-location of planes.
Plane X-location
3 274.2 mm
5 426.6 mm
6 777.0 mm
7 1234.2 mm

Computational analyses by Rodriguez and Cutler
22
showed sensitivity to boundary conditions including
temperature, species composition, and inflow turbulence parameters. They found their solutions to be most sensitive
to the values used for Pr
t
and Sc
t
, and concluded that it is unlikely that constant values for these parameters can be
calibrated to be valid for a wide range of flows. They suggested that the use of models providing for variable Pr
t

and Sc
t
values are necessary to properly simulate the SCHOLAR combustion experiment. Keistler et al.
23
proposed
a model for variable Pr
t
and Sc
t
based on enthalpy and concentration variances, which they used to analyze the
SCHOLAR combustor experiment but comparisons with the data have not shown overall consistency.
Our analysis predicts variable Pr
t
and Sc
t
using the SFM formulation, which includes compressibility
corrections. The facility nozzle from the upstream heater was modeled to provide inflow boundary conditions for
mean and turbulence quantities, to the test section. Near-wall modeling was used for no-slip walls, along with a
constant wall temperature. Early analyses considered the effect of wall temperature and it was found to have a
limited impact on results. This correlates well with the results of Ref. [22], which showed a minor effect from the
temperature of the wall.
B. Grid Sensitivity
A systematic approach was used to model the
SCHOLAR experiment. Three levels of grid refinement
were performed, denoted as Grid A, Grid B, and Grid C
for subsequent levels of refinement. A representative
view of Grid C in the vicinity near the injector is
provided in Figure 13. This domain was decomposed
into an isolator section upstream of the injector, the
injector section containing the angled fuel injector, and a
downstream combustor section. The injector nozzle was
included in a block that extended into the duct. Grid
dimensions are shown in Table 6. The large increase in
injector nozzle grid points was primarily directed at

American Institute of Aeronautics and Astronautics

10

Figure 13. Schematic of SCHOLAR grid near
injector ( every other grid point shown), Plane 5 (X =
426.6mm) shown for reference.

better resolving the nozzle wall boundary layer and turbulence production, with the intent of improving early jet
core mixing predictions.

Table 6. Scholar Grid Dimensions.
Grid A Grid B Grid C
Duct Block 245x95x33 305x113x41
565x145x65
Isolator(black) 209x145x65
Injector(green) 173x145x65
Combustor(purple) 201x145x65
Injector Nozzle Block 134,976 cells 449,196 cells 1,258,353 cells
Total Cells 868,928 1,811,116 6,456,177


Figure 14. SCHOLAR mixing experiment, Wall
pressure grid sensitivity.
Mixing and combustion cases were considered for all
three grids using SFM. Comparisons of lower wall pressure
for the three grids (Figure 14) for the mixing case primarily
show a sharpening of the peaks, particularly downstream of
injection. The general peak locations, however, are basically
the same. Comparisons of results to contours developed from
experimental data are shown in Figure 15 and indicate that
grid refinement helps to better capture the radial extent of the
mixing, but the jet core mixing is being somewhat under
predicted. A three-fold increase in mesh points from Grid B
to Grid C did not significantly improve the core mixing
prediction. This type of trend has been noted in comparisons
to other data sets involving transverse fuel injectors
24
and
work is ongoing to understand the model performance in
these mixing configurations.


(a) mean temperature

(b) mole fraction N
2

Figure 15. SCHOLAR mixing experiment, Comparison of CRAFT predictions to measured data at Plane 5,
grid sensitivity.

American Institute of Aeronautics and Astronautics

11

Wall pressure comparisons to data for all three grids for the combustion case are shown in Figure 16.
Substantial improvement is noted in going from Grid A to Grid C downstream of x =400mm. Grid A was very
coarse stream-wise in this section due to the long combustor, which caused a much more diffusive flow-field than
that of the more refined grids. Contours of mean temperature shown in Figure 17 indicate combustion in Grid A
starting to take place well upstream of the other two grids. However, the amount of combustion taking place is
much less, which coincides with the wall pressure result. The combustion onset location along the top wall for Grid
B and Grid C are in similar locations, providing some consistency in the results. Figure 18 shows mean temperature
and N2 mole fraction contours compared to measured data for all three grids. As in the mixing case, the core jet
mixing is under-predicted, even with the substantial increase in the number of grid points. More combustion
appears to be taking place on Planes 6 and 7 of Grid C compared to the experiment, but the overall contours provide
a better match to data than the coarser grids. These results clearly reveal an improved match to data with grid
refinement going from Grid A to Grid C. However, grid convergence
25
has yet to be realized. Work is progressing
on a formal method of grid convergence, which will be employed in further analyses of this experiment
26
.


(a) lower wall (b) upper wall
Figure 16. SCHOLAR combustion experiment, Wall pressure grid sensitivity.




Figure 17. SCHOLAR combustion experiment, Temperature grid sensitivity.

American Institute of Aeronautics and Astronautics

12



(a) Mean temperature

(b) mole fraction N
2

Figure 18. SCHOLAR combustion experiment, grid-sensitivity comparison.


American Institute of Aeronautics and Astronautics

13


Figure 19. SCHOLAR mixing experiment, wall
pressure predictions vs. data.
C. SCHOLAR Supersonic Mixing Experiment
Having established sensitivities to grid resolution, we
now focus on SFM model comparisons versus those with
constant Pr
t
and Sc
t
, with all calculations performed on
Grid C. The CRAFT CFD

code was used to perform


three-dimensional simulations of this experiment, taking
advantage of planar symmetry on the fuel injector
centerline to reduce the computational expense of the
simulation. The symmetry assumption is consistent with
traditional RANS-based approaches to modeling
symmetrically placed multiple fuel injector arrangements,
and is reasonable for the current comparisons, based on a
review of the experimental data provided in Ref. [20]-[21].
Lower wall pressure predicted with CRAFT CFD

using
variable and constant values of Pr
t
and Sc
t
equal to 0.9 is
compared to measured data from the SCHOLAR mixing
experiment in Figure 19. There is good overall agreement
with the data except for the early peak pressures, which appear to be due to shocks upstream of the injector. These
missed shocks also seem to cause the curves to be shifted downstream of the data. Geometry refinement was
performed to include the upstream facility nozzle to try to capture these shocks, but the two peaks upstream of x =
200mm could not be reproduced in work to date. The same inconsistency to experimental data was observed in
simulations documented by Danehy et al.
20
, suggesting that there may be some physical feature in the experimental
facility that is not being included in relatively simple isolator models. Sensitivity analyses and model refinements
are being considered which may reveal the source of this discrepancy with the data, and its impact on jet mixing.
Results from predictions using both variable and constant values of Pr
t
and Sc
t
equal to 0.9 are compared for
mean temperature and nitrogen mole fraction at a streamwise plane corresponding to x 427mm. As with the wall
pressure, the contours of mean temperature and nitrogen mole fraction between variable and constant Pr
t
and Sc
t
in
Figure 20 are very similar. Contours of local Pr
t
and Sc
t
are shown in Figure 21, exhibiting a range of Pr
t
and Sc
t

from 0.8 to 1.0 in the jet region, indicating that a constant value of 0.9 is an appropriate choice in the downstream
mixing region.
Figure 22 shows CRAFT predictions using variable Pr
t
and Sc
t
compared to the measured data. The general
shape of the jet region is similar between the two, but the extent of jet core mixing is larger in the data. As
mentioned earlier, this discrepancy with data does not seem to be related to grid resolution. While use of constant
values of Pr
t
and Sc
t
for the mixing case has a minimal effect compared to use of the SFM, it has a significant effect
on comparisons to data for the combusting case, as will be discussed.



(a) mean temperature (b) mole fraction N
2

Figure 20. SCHOLAR mixing experiment, Comparison of CRAFT predictions for variable and constant Pr
t

and Sc
t
.


American Institute of Aeronautics and Astronautics

14


(a) turbulent Prandtl number (b) turbulent Schmidt number
Figure 21. SCHOLAR mixing experiment, Predicted Pr
t
and Sc
t
.


(a) mean temperature (b) mole fraction N
2

Figure 22. SCHOLAR mixing experiment, Comparison of CRAFT predictions with measured data.
D. SCHOLAR Supersonic Combustion Experiment
Comparisons of results predicted with CRAFT CFD

for Grid C (using the extended chemistry mechanism and


rates of Conaire, et al.
25
) are made along the symmetry-plane for wall pressure and at four stream-wise planes for
contours of mean temperature and N
2
mole fraction. Figure 23 compares predicted and measured wall pressure
along the upper and lower combustor walls. Three cases are shown in this figure: one using variablePr
t
and Sc
t;
and
the other using a constant Pr
t
equal to 0.9 and Lewis (Le
t
) numbers of 1.0 (Sc
t
=0.9) and 2.0 (Sc
t
=0.45). The
SCHOLAR mixing predictions presented above suggested that constant values of Pr
t
=0.9 and Sc
t
=0.9 would be
reasonable values for analysis of the SCHOLAR combustion problem, since the results between the two were very
similar. For the reacting case, however, predicted wall pressure compared to measured data in Figure 23 shows a
marked difference between the two cases. Here, the Le
t
=1.0 case overpredicts fuel/air mixing with the ignition point
somewhat upstream of the data. The Le
t
=2.0 case grossly overpredicts fuel/air mixing, with the onset of ignition well
upstream of that indicated by the data. The variable Pr
t
/Sc
t
prediction appears to capture the ignition point in much
better agreement with the data, based on the location of the peak pressure spike.


(a) lower wall (b) upper wall
Figure 23. SCHOLAR combustion experiment, wall pressure predictions vs. data.

American Institute of Aeronautics and Astronautics

15

Figure 24 shows contours of temperature, Pr
t
, H
2
O/H
2
mole fraction, and Sc
t
on the symmetry plane for the
variable Pr
t
/Sc
t
case. Along this plane, the combustion begins at both walls near the same X-location, which
correlates to the largest peak in the pressure plots shown in Figure 23. The changes in Pr
t
correspond to sharp
temperature gradients near where the fuel enters the duct and where combustion begins. Pr
t
is greater than 1.0 in
these locations, but is closer to 0.7 near the center of the fuel jet. The plots of H
2
O and Sc
t
reveal Sc
t
values greater
than 0.9 that appear to delay ignition, as was seen in the wall pressure comparison. These figures reveal a continual
variation in Pr
t
and Sc
t
values as the flow evolves downstream. Figure 25 shows Pr
t
and Sc
t
values at the
experimental stream-wise planes, with a variation seen throughout the width of the duct. Contours of Le
t
in Figure
26 reveal the deviation from 1.0 throughout the flow field.






Figure 24. SCHOLAR combustion experiment, symmetry plane contours; temperature,
Pr
t
, H
2
O and H
2
mole fraction, and Sc
t
.

American Institute of Aeronautics and Astronautics

16


(a) Turbulent Prandtl Number (b) Turbulent Schmidt Number
Figure 25. SCHOLAR combustion experiment, Predicted Pr
t
and Sc
t
plotted with mean temperature and
mixture fraction, respectively.




Figure 26. SCHOLAR combustion experiment, contours of Le
t
.

Unlike the mixing problem where constant global values of Pr
t
and Sc
t
were adequate to provide a reasonable
match to the measured data, in the reacting problem these parameters respond more vigorously to sharp
temperature and species gradients in the vicinity of the flame. This result shows the difficulty of selecting constant
values of Pr
t
and Sc
t
which are adequate over the entire domain, for more complex problems, and suggests that their
choice is not intuitive based on the observed behavior of simpler unit problems.

American Institute of Aeronautics and Astronautics

17

Comparisons of predicted to measured contours of mean temperature and nitrogen mole fraction are shown in
Figure 27 for the variable Pr
t
/Sc
t
, Le
t
=1.0 (Pr
t
=0.9, Sc
t
=0.9), and Le
t
=2.0 (Pr
t
=0.9, Sct=0.45) cases. Mean
temperature contours for the variable case show much better comparisons with the data in early portions of the
fuel/air mixing region, with the Le
t
=1.0 case (marked for Sc
t
=0.9) and Le
t
=2.0 case (marked for Sc
t
=0.45)
showing burning to occur prematurely. The constant Pr
t
/Sc
t
cases shown in Figure 27 capture the same behavior
seen in wall pressure, where high combustion temperatures are occurring far upstream compared to the data.
Downstream, the comparisons of the CFD calculations are more reasonable.


(a) mean temperature (b) nitrogen mole fraction
Figure 27. SCHOLAR combustion experiment, comparison of variable and constant Pr
t
and Sc
t

predictions to measured data.
VII. Concluding Remarks
The current status of the SFM was summarized and a building block data base (BBDB) tool for systematic
validation was introduced. Varied basic studies were presented showing the markedly different behavior of scalar
fluctuations in low speed vs. high speed flows, where compressibility-effects introduce complexities into the
modeling. For very basic, non-reacting cases, selection of appropriate values of Pr
t
and Sc
t
can produce adequate
results, but the values to use are not always intuitive. For complex flows, and particular for flows with combustion,
Pr
t
and Sc
t
vary substantially and no constant values are appropriate. We have shown that Pr
t
and Sc
t
change
significantly in flame zones (Burrows and Kurkov and SCHOLAR data) with the changes predicted by the SFM
providing substantive improvements in analyzing complex/combusting data sets. Modeling the effect of
compressibility on scalar fluctuations is a matter of on-going research, particularly as we get into more challenging
supersonic flows with shocks and other complexities.
Acknowledgments
Substantive portions of this effort were performed under an AFRL funded SBIR effort monitored by D. Risha.
References
1
Brinckman, K.W., Kenzakowski, D.C., and Dash, S.M., Progress in Practical Scalar Fluctuation Modeling for High-Speed
Aeropropulsive Flows, AIAA Paper No. AIAA-2005-0508, 43
rd
Aerospace Sciences Meeting and Exhibit, Reno, NV,
J an. 10-13, 2005.
2
Brinckman, K. W., Calhoon, W.H., J r., and Dash, S.M., Scalar Fluctuation Modeling for High-Speed Aeropropulsive
Flows, AIAA Journal, Vol. 45, No. 5, pp 1036-1046, May 2007.
3
Brinckman, K.W., Calhoon, W.H., J r., Mattick, S., Tomes, J . and Dash, S.M.., Scalar Variance Model Validation for High-
Speed Variable Composition Flows AIAA Paper No. AIAA-2006-0715, 44
th
Aerospace Sciences Meeting and Exhibit,
Reno, NV, J an. 9-12, 2006.
4
Calhoon, W.H., J r., Brinckman, K.W., Tomes, J ., Mattick, S. and Dash, S.M.., Scalar Fluctuation and Transport Modeling
for Application to High Speed Reacting Flows AIAA Paper No. AIAA-2006-1452, 44
th
Aerospace Sciences Meeting and
Exhibit, Reno, NV, J an. 9-12, 2006.

American Institute of Aeronautics and Astronautics

18


American Institute of Aeronautics and Astronautics

19
5
Ott, J .D., Kannepalli, C., Brinckman, K.W., and Dash, S.M., Scramjet Propulsive Flowpath Prediction Improvements Using
Recent Modeling Upgrades, Paper No. AIAA-2005-0432, 43
rd
Aerospace Sciences Meeting and Exhibit, Reno, NV, J an.
10-13, 2005.
6
Launder, B.E., Morse, A. Rodi, W., and Spalding, D.B., Prediction of Free Shear Flows, A Comparison of the Performance
of Six Turbulence Models, Free Turbulent Shear Flows, Vol. I Conference Proceedings, NASA Langley Research
Center, Hampton, VA, J uly 20-21, pp. 361-426, 1972.
7
Papp, J .L., and Dash, S.M., Turbulence Model Unification and Assessment for High-Speed Aeropropulsive Flows, AIAA
Paper 2001-0880, 39
th
AIAA, Aerospace Sciences Meeting and Exhibit, Reno, NV, J anuary 8-11, 2001.
8
Sarkar, S., The Pressure-Dilation Correlation in Compressible Flows, Physics of Fluids A, Vol. 4 No. 12, Dec. 1992.
9
Zeman, O., Dilation Dissipation: The Concept and Application in Modeling Compressible Mixing Layers, Physics of
Fluids A, Vol.2, No. 2, Feb. 1990.
10
So. R.M.C., Sarkar, S., Gerodimos, G., and Zhang J ., 1997, A Dissipation Rate Equation for Low-Reynolds-Number and
Near-Wall Turbulence, Theoretical Computational Fluid Dynamics, Vol. 9, pp. 47-63.
11
Sommer, T.P, So, R.M.C, & Zhang, H.S., 1993, Near-Wall Variable-Prandlt Number Turbulence Model for Compressible
Flows, AIAA Journal, Vol. 31 No. 1, pp. 27-35.
12
Kenzakowski, D. C., Papp, J ., and Dash, S. M., Evaluation of Advanced Turbulence Models and Variable Prandtl/Schmidt
Number Methodology for Propulsive Flows, 38
th
AIAA Aerospace Sciences Meeting and Exhibit, AIAA Paper 2000-
0885, Reno, NV, J an. 1013, 2000.
13
Abe, K. and Kondoh, T., A New Turbulence Model for Predicting Fluid Flow and Heat Transfer in Separating and
Reattaching Flows II. Thermal Field Calculations, International J. Heat and Mass Transfer, Vol. 38, No. 8, pp. 1467-
1481, 1995.
14
Calhoon, W.H. Jr., and Kenzakowski, D.C., "Flowfield and Radiation Analysis of Missile Exhaust Plumes Using a Turbulent-
Chemistry Interaction Model," 36th AIAA/ASME/SAE/ASEE J oint Propulsion Conference, Huntsville, AL, AIAA Paper
2000-3388, J ul. 17-19, 2000.
15
Seiner, J .M., Ponton, M.K., J ansen, B.J . & Lagen, N.T., The Effects of Temperature on Supersonic J et Noise Emission,
DGRL/AIAA 92-02-046, DGRL/AIAA 14
th
Aeroacoustics Conference, Aachen, Germany, May 11-14, 1992.
16
Spaid, F.W. and Zukoski, E.E., A study of the interaction of gaseous jets from transverse slots with supersonic external
flows, AIAA Journal, Vol. 6, No. 2, February 1968.
17
Lockwood, F.C. and Moneib, H.A., Fluctuating Temperature Measurements in a Heated Round Free J et, Combustion
Science and Technology, Vol. 22, No. 1-2, pp. 63-81, 1980.
18
Evans, J .S., Schexnayder, C.J ., and Beach, H.J ., Application of a Two-Dimensional Parabolic Computer Program to
Prediction of Turbulent Reacting Flows, NASA Technical Report NASA TP 1169, 1978.
19
Burrrows, M.C. and Kurkov, A.P., Analytical and Experimental Study of Supersonic Combustion of Hydrogen in a Vitiated
Airstream, NASA Technical Report NASA TM X-2828, 1973.
20
P. M. Danehy, S. O'Byrne, A. D. Cutler, and C. G. Rodriguez, "Coherent Anti-Stokes Raman Scattering (CARS) as a Probe
for Supersonic Hydrogen-Fuel/Air Mixing," J ANNAF APS/CS/PSHS/MSS J oint Meeting, Colorado Springs, CO, Dec. 1-
5, 2003.
21
S. O'Byrne, P. M. Danehy, A. D. Cutler, "Dual-Pump CARS Thermometry and Species Concentration Measurements in a
Supersonic Combustor," AIAA-2004-0710, 42nd Aerosciences Meeting and Exhibit, Reno NV, J an 5-8, 2004.
22
Rodriguez, C., and Cutler, A., Computational Simulation of a Supersonic- Combustion Benchmark Experiment, AIAA
Paper 2005-4424, 41st AIAA/ASME/SAE/ASEE J oint Propulsion Conference & Exhibit, Tucson, Arizona 10 - 13 J uly
10-13, 2005
23
Keistler, P.G., Xiao, X., Hassan, H.A., and Cutler, A.D., Simulation of the SCHOLAR Supersonic Combustion
Experiments, AIAA Paper 2007-0835, 45
th
AIAA Aerospace Sciences Meeting & Exhibit, Reno, Nevada 8 - 11 J anuary,
2007.
24
Dash, S.M., Ungewitter, R., Ott, J ., and Papp, J ., Computational Modeling Advances Supporting Hypersonic Scramjet
Design, 14
th
AIAA/AHI Space Planes and Hypersonic Systems and Tech Conf., Canberra, Australia, 6-9 Nov. 2006.
25
Conaire, M.O. Curran, H.J ., Simmie, J .M., Pitz, W.J ., and Westbrook, C.K., A Comprehensive Modeling Study of Hydrogen
Oxidation, International Journal of Chemical Kinetics, Volume 36, Issue 11, 2004. Pages 603-622.
26
Cavallo, P.A., and Sinha, N., Error Quantification for Computational Aerodynamics Using an Error Transport Equation,
Journal of Aircraft, Vol. 44, No. 6, pp. 1954-1963, Nov-Dec 2007.

Das könnte Ihnen auch gefallen