You are on page 1of 17

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

net/publication/270591120

Numerical and Experimental Studies of Flame Stability in a Cavity Stabilized


Hydrocarbon-Fueled Scramjet

Conference Paper · April 2011


DOI: 10.2514/6.2011-2365

CITATIONS READS

7 75

4 authors, including:

Chaitanya Ghodke
Convergent Science Inc., Madison, WI, USA
18 PUBLICATIONS   98 CITATIONS   

SEE PROFILE

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

Roughness-turbulence interactions in oscillatory flows View project

All content following this page was uploaded by Chaitanya Ghodke on 26 May 2015.

The user has requested enhancement of the downloaded file.


17th AIAA International Space Planes and Hypersonic Systems and Technologies Conference AIAA 2011-2365
11 - 14 April 2011, San Francisco, California

Numerical and Experimental Studies of Flame Stability


in a Cavity Stabilized Hydrocarbon-Fueled Scramjet

Chaitanya D. Ghodke∗, Jagannath Pranatharthikaran†, Ghislain J. Retaureau‡


and Suresh Menon§
Computational Combustion Lab.
School of Aerospace Engineering
Georgia Institute of Technology
Atlanta, GA 30332

Flame-holding in a recessed cavity in a Mach 2.5 preheated crossflow is investigated


experimentally and numerically. A sensitivity analysis is conducted on the stability domain
of self-sustained combustion with respect to the operating conditions. The combustor
response is characterized for methane-hydrogen fuel, which is injected from six fuel injectors
located on the floor of the cavity. Small amount of hydrogen addition to methane leads
to significant increase in stability, while reducing the hydrogen beyond lean limit leads to
blowout state. Going from stable combustion to blowout state by reducing the amount of
hydrogen changes the overall heat release and the reaction zone shrinks, which affects the
overall flow structure inside the cavity. Computational studies of chosen test cases using
a compressible large-eddy simulation (LES) approach is carried out using a new turbulent
Artificial neural network (TANN) based filtered chemical rate closure. The differences in
the combustion characteristics for stable and blow-out states are identified and discussed.
The overall results are in good agreement with the experiments and in addition, good
agreement is achieved with the available wall pressure data.

I. Introduction
A significant challenge in the design of an operational SCRAMJET propulsion system is to achieve
efficient fuel-air mixing and combustion within a limited combustor length. The parameters responsible for
mixing are directly affected by chemical and thermal changes taking place in the combustor and there exists
close coupling between fluid dynamics and chemical kinetics in the combustor. In addition to efficient fuel-air
mixing, reliable ignition and flame stabilization over a range of operating Mach numbers are necessary to meet
the operational goals. In particular, for hydrocarbon fueled combustors, flame stabilization is especially of
concern due to the lower propagation speed and heat release of such systems.1 Therefore, various flame-holder
concepts that can provide both efficient flame holding and flame stability are being explored. Designs such
as recessed cavity combustor is one of the promising design that appears to be able to achieve flame-holding
by the generation of subsonic recirculation region that ensures sufficient residence time.2, 3 Additional issues
arise if liquid fuels (e.g., JP-8, JP-10 or kerosene are to be employed4 ) but in this study we focus primarily
on gaseous hydrocarbon fuel such as methane mixed with a small amount of hydrogen.
Past experimental investigations of supersonic cavity flameholders2, 5, 6 have shown that the fuel composi-
tion has a significant impact on flame stability and other parameters such as the injection location and cavity
geometry have also been shown to be critical.7 Methane-fueled cavities usually exhibit a smaller stability
domain than hydrogen- or ethylene-fueled cavities, since methane has a slower chemistry.6 For the purpose
of studying flame stability, estimation of time scales has been used to estimate the Damkohler number8
∗ AIAA Student Member, Graduate Student
† AIAA Student Member, Undergraduate Student
‡ AIAA Student Member, Graduate Student
§ AIAA Associate Fellow, Professor

1 of 16

American
Copyright © 2011 by Chaitanya Ghodke. Published by the American Institute
Institute of Aeronautics
of Aeronautics and Astronautics
and Astronautics, Inc., with permission.
and an empirical model for the blowout limit was demonstrated earlier.9 However, there is insufficient data
over the range of operating conditions and geometrical constraints at this time to determine a generalize
scaling rule for blowout conditions. The studies reported here and elsewhere10 have been carried out with
a particular emphasis on finding blowout limits for a combination of fuel mixtures and for conditions that
have not been studied in the past.
Numerical studies of cavity-stabilized combustion are also sparse. Most of the earlier reported studies
have focused on RANS type modeling, for example, Rodriguez et al.11 simulated the SCHOLAR experiment
using VULCAN CFD code. This same code was used also to study cavity-stabilized ethylene-air flames with
various reduced chemical kinetics mechanisms.12 Another numerical study13 focused on varying the cavity
aft wall angle, the offset ratio and the cavity length for hydrogen injection. The cavity type combustor
was found to enhance mixing and combustion efficiency but at the cost of total pressure loss. Recently,
LES methodology has been used to investigate reacting flow and flame ignition/blowout14 in a a cavity
stabilized combustor and some preliminary results for blowout were reported. This study combined the LES
methodology with an advanced subgrid mixing-combustion model based on the Linear Eddy Mixing (LEM)
sub-grid model within the LES framework (LEM-LES) and this approach was also used recently15 to study
cavity and strut stabilized combustors. In the latter approach a more cost-effective subgrid closure for the
reaction rate kinetics was developed using a turbulent artificial neural network (TANN) approach and was
demonstrated that it can correctly capture the flame dynamics in these type of combustors. The TANN
approach combines the ability of the LEM to capture reaction-diffusion-mixing interactions with finite-rate
kinetics in a look up approach that is both cost effective and accurate in the regimes of interest. Earlier
studies16, 17 have demonstrated the ability of this approach in more canonical flows.
In this paper, we report on experimental and numerical efforts to study cavity stabilized combustion
in a Mach 2.5 cross-stream. Experiments are conducted for several methane-hydrogen fuel blends, while
numerical studies are carried out for only few chosen experimental cases. To ensure proper matching of the
experiments, the entire test facility is modelled in the LES. Comparison with data where available is carried
out and the LES results are used to explain the flow and flame dynamics in the experimental cases for both
stable and blowout conditions.

II. Experimental Study


II.A. Experimental Setup
The Georgia Tech supersonic test facility is shown in Fig. 1. A blow down system and a heater provide a
non-vitiated heated primary air flow. The stagnation pressure can be adjusted during the experiment within
a 101 - 2169 kPa range, whereas the stagnation temperature can reach up to 750 K depending on the mass
flowrate. Although these conditions are relevant, the stagnation temperature is much lower than the actual
flight condition for the Mach 2.5 flow used here. However, these conditions are still relevant to show the
overall features of flame stability and blowoff.
The test section is made of stainless steel and is depicted in Fig. 2. The nozzle provides a Mach 2.5 flow
into a 31.8 mm × 63.5 mm rectangular cross-section test section. The cavity used for these experiments is
D= 25.4 mm deep, and L= 97.5 (L/D= 3.84). A diverging ceiling starts 127 mm from the leading edge of
the cavity with a angle of 2.5o to allow for heat release and thermal expansion effect from the combustion.
The exit of the test section is at atmospheric pressure.
The fuel injection system ensures the delivery of a CH4 -H2 blend at 297 K through an array of six injectors
equally spaced along the spanwise direction and located on the floor of the cavity, 6.4 mm downstream the
leading step. Each injector is 2.3 mm in diameter. The fuel mixture composition is adjustable using two
mass flow controllers (1000 SLPM - 100 SLPM). A 6000 V spark is also present at the bottom of the cavity.
Pressure is measured at various wall locations to track shock patterns and to monitor the combustor
startup and operation (P1−7 in Fig. 2) with an accuracy of 0.5 to 1 kPa. The cavity is also instrumented
with a pressure and a temperature port located on the bottom wall 50.8 mm from the leading step. For
temperature measurements, thermocouple probes are present in the settling tank, the fuel line, the cavity
and at the ceiling of the test section. The crossflow temperature Tcf is estimated from T1 which is located
next to P1 in the spanwise direction such that P1 -T1 and Pcav -Tcav are mirror image with respect to the
channel center plane. All thermocouple are K-type and are directly mounted on a cold junction compensator.
The accuracy of all K-type thermocouples is ± 0.75% of the measured temperature in Celsius.

2 of 16

American Institute of Aeronautics and Astronautics


Figure 1. Facility Schematic. 1) Air storage tank (9755 scm at 20 MPa). 2) Shut off valve. 3) Air flow controller.
4) Heater (Burner up to 810 K). 5) Heated pipe. 6) Manual shut off valves. 7) Insulated pipe (65 mm glass wool
and thermally reflecting sleeve). 8) Settling tank. 9) Expansion joint (thermally insulated). 10) Flow straightener
(thermally insulated). 11) Test section. 12) Exhaust. A) Methane storage tank (1675 scm at 34.5 MPa). B) Fuel
cylinders with pressure regulators and shut off valves. C) Methane shut off valve and flow controller. D) Fuel control
and injection system.

Figure 2. Test section cut view, L/D = 3.84 shown (dimensions are in mm).

3 of 16

American Institute of Aeronautics and Astronautics


II.B. Ignition-Blowout Procedure
Ignition is usually performed at a favorable fuel flowrates (0.6 g/s) and hydrogen dominant fuel mixtures
(90% hydrogen and 10% of methane in mass proportion), typically far off from the blowout conditions. A
sudden temperature rise is recorded by the thermocouple located on the cavity floor and then the igniter
and the additional hydrogen supply are switched off. If the combustion continues without the ignitor it then
defines the self-sustained combustion which studied for stability analysis. The combustor is maintained at
this regime until wall steady state temperature is reached, which takes about 10 seconds. Within a couple
of seconds, the fuel mixture is then adjusted to a gross mixture composition target. However, if the mixture
changes rapidly, the controller readings do not correspond to the actual mixture being injected since there
is a time for the fluid to travel from the mass flow controller to the injectors. In addition, the combustor
takes some time to stabilize under constant input, and this is accounted in the post processing of the data
as well. Therefore in the next step, a slow tuning phase is necessary to ensure that the mixture composition
injected in the cavity fairly coincides with the controller readings. Blowout is studied at constant stagnation
pressure and temperature such that the combustion may cease due to lack of fuel or oxidizer. Figure 3(a)
shows a typical ignition-blowout procedure as a function of the fuel mixture composition.
To provide some interpretation, the overall mixture is represented in a three dimensional space where
a given regime is a point whose coordinates are the mass flow rates of methane, hydrogen and air. Figure
3(b) shows the experimental data in terms of trajectories in the mixture space. These are the paths that
satisfy ignition and burning so it is possible to estimate the size of the stable combustion domain. All the
trajectories terminate in blowout points. The blowout limit draws a region that separates stable combustion
from the mixtures for which no combustion is sustainable. The stable combustion domain forms a volume
in which lies all the mixture compositions that resulted in stable combustion in the cavity. This volume is
expected to be a function of the preheat temperature, the fuel injection location and the cavity aspect ratio.

(a) Ignition - Blowout procedure in the mixture space. (b) Stable domain and blowout region in the mixture space:
Mixture trajectories (grey lines) and blowout events (black
symbols).

Figure 3. Construction of the experimental blowout region : (a) depicts a single orbit in the mixture space whereas
(b) is the collection of all orbits (burn data). The end points form the blowout region.

An efficient algorithm is used to post process the data in which the temperature measured in the cavity
is used to detect the ignition and blowout. This process is described in a previous paper.10

II.C. Results and Discussion


Although many cases were studied earlier,10 the test cases (with and without preheat) we are focussing on
are summarized in Table 1.
Figure 4 shows the flame location during stable combustion and near lean blowout. As the fuel rate
is decreased, the flame shifts toward the injector region. This behavior has already been observed in the

4 of 16

American Institute of Aeronautics and Astronautics


Fuel Flowrate L/D Pcf Tcf
Case 1 0-1.3 (g/s) 3.84 40-65 kPa 135 K
Case 2 0-1.3 (g/s) 3.84 40-65 kPa 248 K
Table 1. Test matrix

past9, 18 and suggests that the flow and the combustion mechanisms depend on the fueling rate. Henceforth,
the physics involved in rich and lean blowout differ and the results reported in this study are specific to the
combustion regime.

(a) Stable combustion (mf = 0.48 g/s). (b) Lean combustion near blowout (mf = 0.31 g/s). The
flame is located near the injector region.

Figure 4. Flame location as function of the fueling rate.

The blowout data points are located on a map with respect to the fuel mixture in Fig. 5(a) for both
Case 1 and 2. The blowout data is represented as a cloud of points which separates the stable combustion
from the non-burning region (close to the ṁCH4 axis). Even in small quantities, addition of hydrogen in
the fuel blend mixture is advantageous to increase the stability of the combustor. The air-methane flame is
stabilized with small hydrogen addition, typically as low as 2 % of the fuel mass flowrate.
As the crossflow static pressure is lowered, the combustion mechanism is maintained in less favorable
conditions: the data dispersion increases and the blowouts become less repeatable. Fuel-lean (ṁf < 0.4 g/s)
and fuel-rich regions (ṁf > 0.7 g/s) are more affected by this phenomenon. Eventually, no combustion is
obtained below 40 kPa of cavity static pressure within the tested conditions, such that the static pressure
draws a lower bound below which no combustion is sustainable. Higher hydrogen concentrations show
that the lack of pressure can be compensated by higher flame speed and mixture diffusivity in the sense
of non-premixed time scale of Driscoll et al.8 Further observations shows that pressure may also shift the
stochiometry of the mixture in the combustor toward the lean limit since the air entrained in the supersonic
shear layer is more dense as predicted by the work of Dimokatis.19, 20
Impact of the supersonic crossflow temperature on the combustion stability is also investigated (Case 2 of
Table 1). The blowout fuel maps conserves the same trends however, the blowout limit trend is shifted down
to lower hydrogen flow rates (as low as 200 parts of methane for 1 part of hydrogen) leading in an increase
of the stable domain. In the preheated case, the amount of energy spent in raising the incoming mixture
to its flammability limits is reduced. The extension of the stable domain may also be due to better flame
anchoring provided by the higher unburnt gas temperature in the lift off region described by Rasmussen and
al.3 In addition for Case 2, the wall heat losses are reduced since the structure is heated and the temperature
gradient may be therefore reduced.
The symptoms of having a low crossflow pressure and temperature are the poor repeatability (high
dispersion of the blowout data points) and a hydrogen increase in the fuel mixture. These results shows that
adjustable fueling composition can play a role in maintaining combustion and extending the stable domain
by offering more flexible heat content.
Further studies on cavity blowout lead to the establishment of an empirical model by Rasmussen et al.9
Lean blowout limits with floor injection follows Φ = 0.0028Da−0.8 at best correlation. The Damkohler number
is given by the ratio of characteristic time scale between the flow and the chemical reaction.9

5 of 16

American Institute of Aeronautics and Astronautics


D
ucf
Da =     (1)
αo 101325 P a 300 K
2
SLo Pcf Tcf

The overall mixture properties such as stoichiometric flame speed and thermal diffusivity are calculated
using Cantera. The equivalence ratio expressed from the characteristic air mass flowrate.3

ṁ∗ = ρcf ucf LW (2)


where L and W are the length and the width of the cavity, respectively. Figure Fig. 5 depicts the blowout
data with respect to the equivalence ratio and the Damkohler number. When approaching lean blowout,
the flame region is located upstream near the injectors.21 This criterion is used to separate lean and rich
blowout events. The overall data agrees well with the model proposed by Rasmussen et al.9 as seen in Fig.
5(b) in terms of order of magnitude, but some deviation is observed. Equations 1 and 2 and can be used to
explain the observed trends when changing the crossflow temperature, cavity length and fuel mixture. The
effect of crossflow temperature on the blowout margin can be seen from Eq. 1. The crossflow velocity in Eq.
1 is approximated as

ucf = Mcf (γRTcf )1/2 (3)


Assuming a perfect gas and neglecting dissociation between low and high crossflow temperature, the
T 1/2
flow time scales with the inverse of the square root of the crossflow temperature and Da ∝ ucf cf
∝ Tcf .
Consequently, preheat produces a shift of the data toward the right side of the graph, as observed. In
−1/2
addition, the characteristic air mass flowrate decreases with temperature as ṁ∗ ∝ Tcf . This may explain
why richer blowout mixtures are found at 285 K crossflow temperature. Calculations also reveals that the
stoichiometric flame speed increases with the thermal diffusivity, which produces additional dispersion along
the Da -axis.

(a) Fuel mixture composition at blowout: the stable do- (b) Equivalence ratio as a function of Damkohler number
main is located on the right side of the cloud of points. at lean blowout: comparison with blowout data from other
facilities,9 the legend reads as Facility/Fuel/Ramp Type.

Figure 5. Experimental blowout data.

6 of 16

American Institute of Aeronautics and Astronautics


III. Numerical Study
IV. Mathematical Formulation
IV.A. LES Filtered Equations
The LES governing equations are obtained by applying a spatial filter based on grid size, ∆. The Favré-
filtered variable defined as: fe = ρf /ρ, where ρ represents the local fluid density and f¯ denotes the non-density
weighted filtered variable f . The compressible Navier-Stokes equations for the mass, momentum, total energy
and species densities conservation in generalized co-ordinates are written as

∂ρ ∂ρuei
+ =0 (4)
∂t ∂xi
∂ρuei ∂  sgs 
+ ρuei uej + pδij − τij + τij =0 (5)
∂t ∂xj
∂ρEe ∂ h e  i
+ ρE + p uei + qi − uej τij + Hisgs + σisgs = 0 (6)
∂t ∂xi
∂ρY
fk ∂ h f 
sgs sgs
i
+ ρ Yk uei + Y
fk V
g i,k + Yi,k + θi,k = ω˙k k = 1, ..., Ns (7)
∂t ∂xi
where variables with overbar denotes spatially filtered quantities, variables with tilde denote Favre averaged
quantities and variables with superscript sgs represent unclosed subgrid scale terms. In the above equations,
ρ is the density, ui is the Cartesian velocity component, p is the pressure, τ is the shear stress, E is the total
energy, q is the energy flux due to thermal conduction and species diffusion, Yk is the species mass fraction
and ω˙k is the reaction rate. Pressure, p is evaluated using the filtered equation of state, p = ρR̃T̃ + ρT sgs .
2

The viscous stress, τij in Eq. 5 is evaluated from τ ij = 2µSij + µd − 3 µ Skk δij where, δij is the Kronecker
delta function, Sij = 1/2(∂ui /∂xj + ∂uj /∂xi ) is the resolved rate of strain, µ is the dynamic viscosity and
µd is the bulk viscosity evaluated using the formulation given in Ern and Giovangigli.22 The thermal flux,
qi in Eq. 6 is given by q̄i = −κ̄ ∂x∂Te + ρ̄ PNs h̄ Ye Ve where Fick’s law for species diffusion and Fourier’s law
i k=1 k k i,k
for thermal conduction are assumed.
To close the sgs terms, τ sgs , H sgs , σ sgs and Y sgs in Eqs. 5-7, a transport equation for subgrid Kinetic
Energy is solved:
h i
∂ sgs ∂ sgs ∂ ∂ksgs ρνt R
e ∂ Te
∂t ρ k + ∂xi (ρ u
ei k ) = ∂xi (ρνt + µ) ∂xi + P rt∂xi
 
sgs 2 ρS ksgs
2 
sgs ∂ e
uj (ksgs )3/2
 (8)
− 1 + αpd Mt e
Dksgs τij ∂xi + ρc ∆

23
The coefficients, cν and c in the above
√ equation are evaluated using a dynamic procedure. Eddy viscosity,
νt computed from k sgs as νt = Cν k sgs ∆ is then used in conjunction with the eddy viscosity hypothesis
and the gradient closure hypothesis to close the subgrid viscous stress, τ sgs and subgrid species flux Y sgs
respectively. In addition, the turbulent Prandtl number, P rt used to close the energy equation, is also com-
puted dynamically and locally.23 The closure procedure and a comprehensive discussion of the formulation
used in this work can be found in Genin and Menon23 and skipped here for brevity. The computation of ω̇
is discussed further below.

IV.B. ANN Based Chemical Kinetics Modeling


The cost of solving finite-rate kinetics in LES is very significant with or without the LEM closure. This is
primarily due to the stiffness of the chemical mechanism. Methods based on mixture fraction and flamelet
modeling have been developed to avoid this issue but in situations where there are multiple fuels and
combustion ranges from non-premixed to premixed state as in the cavity, such an approach does not make
much sense. Finite-rate kinetics are needed but their cost overhead needs to be reduced. An alternative
strategy involves solving the LES equations represented by Eqs. 4-8 and evaluating the filtered reaction rate
from an Artificial Neural Network (ANN) based approach.16, 17 This involves the creation of a look-up table
that contains the (filtered) thermo-chemical state-space accessed by the reacting system. The look-up table

7 of 16

American Institute of Aeronautics and Astronautics


is created offline by simulating a 1D non-premixed flame configuration using a standalone 1D LEM solver
in the manner considered by Sen and co-workers.17 Consequently, the ANN, essentially an interpolation
scheme that relates to a set of inputs to a set of outputs through a network of interprocessing elements,
is then trained based on this tabulated information. Finally, ANN is used online with the LES solver to
determine the reaction rates for thermo-chemical states that the ANN was not necessarily trained for. The
aspect of unsteady interaction of chemical kinetics with turbulence is inherent in the training data due to
the multiscale LEM methodology discussed above and has been found to be superior to other ANN training
approaches based on laminar flame data.24 Overall, the ANN-LES approach is found to yield significant cost
and memory savings with acceptable accuracy.15–17, 24
ANN has been used within LEMLES simulations to cut down the computational cost associated with
solving a stiff ODE to compute the subgrid reaction rates for both premixed and non-premixed canonical
flame configurations.15, 17, 24 In this study, the ANN strategy16 is implemented by constructing a training
table based on the filtered species mass fractions, filtered species gradients, filtered temperature and subgrid
Reynolds number to compute the filtered rate as ω̇ ¯ i = f (Y˜k , T̃ , Re∆ , ∂ Y˜k ). The filtered quantities are
∂xi
computed from the LEM domain at the time of table-preparation by assuming a filter size determined by the
LES grid. The tabulated data span thermo-chemical states from multiple realizations per turbulent Reynolds
number for a range of turbulent Reynolds numbers accessed by the LES simulation. 85% of the standalone
LEM data is used for training the net. Following training, the ANN is tested within the standalone code
for prediction of rates for inputs culled from the untrained state-space to check for accuracy and robustness
of training as shown in Appendix, Fig. 17. The end result is an ANN that yields the filtered reaction rate
online when fed the instantaneous filtered thermo-chemical data from a LES simulation. This methodology
is henceforth called the Turbulent-ANN LES (TANN-LES) approach.

V. Numerical Approach
A block structured finite-volume methodology23 that is second order accurate in space and time is used
to solve the LES governing equations described in Section IV. The evaluation of the fluxes in finite-volume
algorithm is performed using a hybrid methodology developed for the resolution of high speed turbulent flow
environment characterized by shock-turbulence interactions. The algorithm has been validated and applied
for supersonic applications elsewhere23 and is implemented in a multiblock parallel framework using the
Message Passing Interface (MPI) library. The time-step being used for the time integration of the governing
equations is determined from stability consideration for the advection and the diffusion equations.

V..1. Computational Configuration


LES simulations of the Georgia Tech supersonic test facility with Mach 2.5 inlet nozzle are carried out.
Total number of computational cells is approximately 4.5 million with clustering of grid points in regions of
high gradients such as the near-wall region, shear layer and the fuel injectors. The smallest cell in the wall
normal direction is 0.254 mm. Guided by experiments, nominal inlet conditions given in Table 2 are used.
Supersonic boundary conditions are imposed at inflow and outflow. No-slip wall boundary condition is used
at the top, bottom and side walls. The blended fuel mixture of methane and hydrogen is injected at the
bottom of the cavity at P = 130 KPa and T = 291 K. Combinations of methane and hydrogen mass flow
rates corresponding to one stable combustion and one near blowout cases are used 1.
Table 2. Mach-2.5 flow conditions used for analysis

Case Inlet P0 (KPa) Inlet T0 (K) ṁCH4 (g/s) ṁH2 (g/s)


Stable 859.2 530 0.68 0.05
Blowout 801 529 0.63 0.03

In this study, a reduced four-step Methane-Hydrogen mechanism25 is employed:

CH4 + 2H + H2 O → 4H2 + CO (9)


H2 O + CO → H2 + CO2 (10)

8 of 16

American Institute of Aeronautics and Astronautics


2H + M → H2 + M (11)
3H2 + O2 → 2H + 2H2 O (12)

where M is a radical. The TANN is trained on this kinetics for the range of turbulent conditions in the
combustor.

VI. Results and Discussion


Distinctly different flow features and reaction zone structures are observed for stable and blow-out cases.

(a) Temperature field (b) Product (H2 O) field

Figure 6. Time-averaged flowfield at the cavity centerplane for stable combustion case

Figure 6 shows time averaged temperature and H2 O field in the cavity for the stable combustion case.
In this case, there exists an efficient fuel/oxygen mixing region that sustains stable combustion with hot
products recirculating in the cavity, maintaining high temperature. Hence, overall cavity shows region
of high temperature and significant amount of products. Figure 7(a) and Fig. 7(b) show instantaneous
temperature field at cavity centerplane and at 3 spanwise locations (x = 0.18, 0.23, 0.26 m), respectively.
Combustion takes place in most of the cavity region where both fuels are being mixed with the entrapped
oxidizer. Peak temperature occurs at the corner upstream of the injectors but is also well distributed in the
entire combustor. Spanwise temperature field in Fig. 7(b) shows 3D wrinkled structures with local regions
of high temperature. When compared to LEMLES approach,14 turbulence flame interactions appear to be
well captured by TANN-LES approach and in computationally much affordable way.
To understand flow features inside the cavity in stable combustion case, velocity vectors are plotted and
are shown in Fig. 8. It can be clearly seen that a large vortical structure forms and spans over almost entire
cavity length (shown by red box). This vortical flow helps hot products of combustion to recirculate inside
the cavity and preheat oncoming fresh fuel and oxygen. This subsonic recirculation region provides more
residence time for fuel/air to get well mixed and react with each other. This appears to be the primary
mechanism of flame holding for the stable combustion case.
Although fuel is injected into the cavity and therefore, this combustion problem could be classified to
be non-premixed combustion, due to product recirculation and variable mixedness, the combustion process
is much more complex than a simple diffusion flame. To delineate between premixed and non-premixed
∇YF ·∇YO
flame fronts and analyze the flame structure, the flame index26 given by, Flame Index (F.I.) = |∇Y F ·∇YO |
is calculated; where F denotes the fuel and O being the oxidizer. F.I. is positive for premixed flame and
negative for diffusion flame. In premixed region, the gradients of fuel and oxidizer are aligned. So the positive
values of F.I. are seen due to rapid consumption of the reactants across the flame. In non-premixed flames
the gradients of fuel and oxidizer oppose each other, and hence F.I. is a negative value. Fig. 9(a) shows
flame index at stream-wise location for stable combustion case indicating both premixed and nonpremixed
burning occurs inside the cavity. Negative values of F.I. are mostly observed near the shear layer and near
the injectors where diffusion flame is present, while positive values of F.I. showing premixed burning are
observed mostly in the aft cavity portion. As discussed earlier, presence of a large vortical structure inside
the cavity increases local flow time and enhances fuel/air mixing process and this is confirmed by the positive
values of F.I. in this region. In this premixed zone, CH4 / air could burn by itself even there is very little
or no H2 present.

9 of 16

American Institute of Aeronautics and Astronautics


(a) Temperature field at cavity centerplane

(b) Temperature field at 3 spanwise locations: x = 0.18, 0.23, 0.26 m from left to right

Figure 7. Instantaneous flowfield for stable combustion case

Figure 8. Velocity vectors showing recirculation region inside cavity in stable combustion case (zoomed in view of
cavity)

(a) Flame index at centerplane (b) Stoichiometric mixture fraction overlaid on tempera-
ture field

Figure 9. Flame structure in stable combustion case

10 of 16

American Institute of Aeronautics and Astronautics


Figure 9(b) shows the stoichiometric mixture fraction contour lines overlaid on temperature field showing
presence of diffusion flame surface near the injector and the shear layer. As it can be seen, reaction zone spans
over overall cavity in stable combustion case. In summary, large recirculation region of hot temperature field
and presence of premixed fuel/air mixture inside the cavity help to sustain the stable combustion process.
A near blowout case is designed by reducing amount of hydrogen injected to see its effects on combustion
characteristics. As can be seen in Fig. 10(a), hot region of temperature is now mostly seen near the injector
location only. Also, unlike stable combustion case, not significant amount of combustion products is present
inside the overall cavity. Figure 10(b) shows time averaged H2 O field inside the cavity for near blowout case
showing presence of products near injector location only. There is still some high temperature and more
products in the upstream cavity corner but this region is essentially trapped by the injector flow and does
not contribute significantly to the combustion process in the rest of the cavity.

(a) Temperature field (b) Product (H2 O) field

Figure 10. Time-averaged flowfield for near blowout case

Figure 11 shows zoomed in view of the time-averaged velocity vectors inside the cavity in near blowout
case. When compared to the stable case there are substantial differences. There is no single large recirculation
region present inside the cavity (as opposed to stable combustion case). Instead, there are small local regions
present (shown by red box) in the aft cavity. As a result, there is no efficient feedback mechanism present,
which can recirculate hot products of combustion inside the cavity and preheat oncoming fuel/air mixture.
In the near blowout case, as less hydrogen is blended with methane, there is no enough heat release for
thermal expansion of flow inside the cavity. The reduced temperature and thermal expansion effect seems to
breakup the recirculation in the cavity. Small local vortical structures can recirculate only locally and thus,
seem to be inefficient in capturing the overall mixing process.

Figure 11. Velocity vectors showing recirculation regions inside cavity - near blowout case (zommed in view of cavity)

Flame index in Fig. 12(a) shows diffusion type burning (black color) near the injector location. Due to
absence of strong recirculation current, there is hardly any premixed region seen in this case. Stoichiometric
mixture fraction contours overlaid on temperature field in Fig. 12(b) show primarily diffusion flame surface.
It can be clearly seen that, flame is mostly seen near injector location. In effect, reaction zone is shrunk and
this maybe the cause of the blowout event. Figure 13 shows CH4 field in near blowout case. As it can be
seen, lot of CH4 is left and is escaping out of cavity without getting mixed with oxidizer inside cavity. This
could be due to absence of strong recirculation current in the near blowout case.
Further analysis of the thermal expansion effect and the sensitivity of the combustion process will be
carried out in the future when more cases are simulated and when other fuel-air mixtures are considered.

11 of 16

American Institute of Aeronautics and Astronautics


(a) Flame index at centerplane (b) Stoichiometric mixture fraction plotted over tempera-
ture field

Figure 12. Flame structure in near blowout case

Figure 13. CH4 contours in blowout case

For now, the current studies agree with experimental observations that the flame structure and the hot
temperature region changes significantly from stable to unstable combustion. Finally, an estimate of the
Damkohler number based on local properties is carried out (note that this is a local estimate as opposed to
a global estimate in the experimental case). Equivalent to Fig. 5(b), scatter plot of Damkohler number as
a function of local equivalence ratio is plotted in Fig. 14 along with relevant global data. There are some
clear differences between the local Da variation and the global estimate. Further research is still needed to
understand this discrepancy.

Figure 14. Equivalence ratio as a function of Damkohler number - numerical study

Figure 15 shows the top wall pressure comparison with available experimental data. It can be seen that,
locations of shocks and peak pressure inside the combustor are predicted reasonably well. Location of leading
edge shock (x ≈ 0.22 m), expansion (x ≈ 0.26 m) followed by shock at (x ≈ 0.32 m) are well predicted.
Multiple reflections off the wall (x ≈ 0.34 m, 0.37 m and 0.42 m) are also captured well.

12 of 16

American Institute of Aeronautics and Astronautics


Figure 15. Shock structure and top wall pressure comparison with experimental data

VI..2. Computational Time Savings


Computational time required for TANN-LES in this study is compared with LEMLES approach used earlier14
shown in Table 3. Substantial reduction in the computational cost is obtained using TANN-LES approach.
Species equation is solved on the LEM level in LEMLES approach, while TANN-LES solves it on the LES
level. Therefore, speed-up obtained using TANN-LES approach is much more than that can be obtained
using LEMLES. Compared to the LEMLES, TANN-LES provides speed-up of around 42 times. Speed-up
obtained could be even more if detailed mechanism is considered.
Table 3. Speed-up obtained for 4 steps methane mechanism

Method Species Equation Time/(step×cell) Speed-up


TANN-LES LES Level 2.9476×10−2 42
LEMLES LEM Level (LES/12) 1.2352 1

VII. Conclusions
This paper discusses a combined experimental and numerical study of ignition and blow-out limits of
methane-hydrogen fuel blend in high speed flow environment. Experiments are conducted to study combus-
tion ignition and blow-out limit in a Mach 2.5 crossflow in a facility developed at Georgia Tech. Results
show that addition of even a small amount of hydrogen greatly influences the reaction time, altering the
combustion mechanism in a more favorable manner. The blowout data collected during experiments is com-
pared to the data from other facilities. Subsequently, LES of one stable and one blowout case guided by
experiments are carried out. ANN based filtered chemical rates modeling (Turbulent-ANN) approach for
turbulent combustion is employed. The flow structures observed in stable combustion and blow-out cases
show distinct differences. In stable combustion, a large vortical flow is formed in the aft region of the cavity

13 of 16

American Institute of Aeronautics and Astronautics


that provides an effective mechanism to transport the hot products and to enhance mixing of the fuel-air
mixture. Combustion occurs in non-premixed mode in the injector shear layers and the cavity shear layer,
while majority of the cavity is burnign in a premixed mode. On the other hand, small localized vortical
flow structures are seen in the near-blowout case that do not seem to be as efficient in mixing the hot prod-
ucts and/or in transporting the premixed mixture. Reaction zone is shrunk in the blowout case and only
non-premixed flame is mostly seen near injector location. These results appear to be in excellent agreement
with experimental observations. Predictions of wall pressures also demonstrate good agreement with exper-
imental data. Study of possible correlation between Damkohler number and local equivalence ratio based
on local flow properties is performed and some discrepancies are seen but this is not yet resolved. These
results demonstrate that a closely collaborated experimental-numerical study can be used to understand and
explain observations that otherwise cannot be fully resolved due to difficulty in in-situ measurements.

VIII. Appendix
VIII.A. Combustion Regimes
In order to determine different combustion regimes and understand the turbulent flame structure a Borghi
diagram modified by Pitsch27 for LES applications is constructed based on premixed F.I. values. The analysis
requires total RMS velocity (u0∆ ), LES filter size (∆), local laminar flame thickness (LF ) and local laminar
flame speed (SL ). The RMS velocity and the LES filter width (same as grid size) can be evaluated using
the data obtained from the simulations. However, the laminar flame speed and the flame thickness will
also depend on amount of products present at particular location and they need to be estimated locally
everywhere inside the combustor. Using the flame speed and the flame thickness, Karlovitz number (Ka)
defined for LES27 is evaluated. Figure 16(a) shows different flame regimes seen in the present study. Strong
variation of Ka number from corrugated flamelet to broken reaction zone is observed. For Ka < 1 turbulent
motions wrinkle the flame front causing convolution leading to formation of pockets of fresh and burnt
gases. For 1 > Ka > 100, turbulent motions enter and modify flame preheat zone. For Ka > 100, both
diffusion and reaction zone are affected by turbulent motions. Figure 16 shows probability distribution of
flame regimes inside the cavity combustor. Injector plane location is at x ≈ 0.006 m (x = 0 is at cavity LE).
It can be seen that the probability of broken zone and thin reaction zone is more near the injector location.
This could be the effect of high level of turbulence due to floor injection. Thin reaction zone dominates in
the overall cavity, whereas flamelet type burning is mostly seen at aft portion of the cavity. It is emphasized
that this analysis only looks that the premixed regions based on the F.I. values.

(a) Borghi diagram for LES (b) PDF of flame regimes inside the combustor

Figure 16. Flame regimes and their probability inside the combustor

14 of 16

American Institute of Aeronautics and Astronautics


VIII.B. Validation of TANN-LES Approach
Figure 17 shows the correlation curve obtained for the 10/8/4 TANN. It can be seen that, the correlations
for all the species collapse on top of each other indicating the overall accuracy during the training phase.
Predictions are highly correlated with the target value, except for some points. Considering that the plot
shows almost 90,000 points, there are only few points outside the target values and they do not affect the
overall performance.

Figure 17. Correlation between the target value and TANN prediction for all species

IX. Acknowledgements
This work is supported in part by the Air Force Office of Scientific Research. The support at the Depart-
ment of Defence (DOD) High Performance Computing Center (HPC) of Army Research Laboratory(ARL),
NAVY and Engineer Research and Development Center (ERDC) for simulations is acknowledged.

15 of 16

American Institute of Aeronautics and Astronautics


References
1 K. H. Yu, K. J. Wilson, and K. C. Schadow. Effect of flame-holding cavities on supersonic combustion performance.

Journal of Propulsion and Power, 17(6):1287–1295, 2001.


2 M. R. Gruber, R. A. Baurle, T. Mathur, and K. Y. Hsu. Fundamental studies of cavity-based flameholder concepts for

supersonic combustors. Journal of Propulsion and Power, 17(1):146–153, 2001.


3 C. C. Rasmussen, J. F. Driscoll, C. D. Carter, M. R. Gruber, J. M. Donbar, and K. Y. Hsu. Stability limits of cavity-

stabilized flames in supersonic flow. Proceedings of the Combustion Institute, 30:2825–2833, 2005.
4 V. A. Vinogradov, S. A. Kobigsky, and M. D. Petrov. Experimental investigation of kerosene fuel combustion in supersonic

flow. Journal of Propulsion and Power, 11(1):130–134, 1995.


5 T. Mathur, M. R. Gruber, K. Jackson, J. Donbar, W. Donaldson, T. Jackson, and F. Billig. Supersonic combustion

experiments with a cavity-based fuel injector. Journal of Propulsion and Power, 17(6):1305–1312, 2001.
6 M. R. Gruber, J. M. Donbar, C. D. Carter, and K. Y. Hsu. Mixing and combustion studies using cavity-based flameholders

in a supersonic flow. Journal of Propulsion and Power, 20(5):769–778, 2004.


7 C. C. Rasmussen, J. F. Driscoll, C. D. Carter, and K. Y. Hsu. Characteristics of cavity-stabilized flames in a supersonic

flow. Journal of Propulsion and Power, 21(4):765–768, 2005.


8 J. F. Driscoll and C. C. Rasmussen. Correlation and analysis of blowout limits of flames in high-speed airflows. Journal

of Propulsion and Power, 21(6):454–457, 2005.


9 C. C. Rasmussen and J. F. Driscoll. Blowout limits of flames in high-speed airflows: Critical damkohler number. AIAA

2008-4571, 44th AIAA/ASME/SAE/ASEE Joint Propulsion Conference and Exhibit, 21 - 23 July 2008, Hartford, CT, 2008.
10 Kovitch S. B. Verma S. Retaureau, G. J. and S. Menon. Experimental studies of cavity flame-holding in a Mach 2.5 cross

flow. AIAA-2009-0810, 47th AIAA Aerospace Sciences Meeting, 5 - 8 Jan 2009, Orlando, Florida, 2009.
11 C. G. Rodriguez and A. D. Cutler. Computational simulation of supersonic-combustion benchmark experiment. AIAA

2005-4424, 41st AIAA/ASME/SAE/ASEE Joint Propulsion Conference and Exhibit, July 10 - 13, 2005, Tucson, Arizona,
2005.
12 J. Liu, C.-J. Tam, T. Lu, and C. K. Law. Simulations of cavity-stabilized flames in supersonic flows using reduced

chemical kinetics mechanisms. AIAA 2006-4863, 42nd AIAA/ASME/SAE/ASEE Joint Propulsion Conference & Exhibit,
Sacramento, California, 2006.
13 Kyung Moo Kim, Seung Wook Baek, and Cho Young Han. Numerical study on supersonic combustion with cavity-based

fuel injection. International J. Heat and Mass Transfer, 47, 1991.


14 J.J. Choi, C. Ghodke, and S. Menon. Large Eddy Simulation of cavity flame-holding in a Mach 2.5 cross flow. AIAA

2010-414, 48th AIAA Aerospace Science Meeting, Orlando, FL, 2010.


15 C. Ghodke, J.J. Choi, and S. Menon. Large Eddy Simulation of supersonic combustion in a cavity-strut flameholder.

AIAA 2011-323, 49th AIAA Aerospace Science Meeting, Orlando, FL, 2011.
16 B. Sen and S. Menon. Artificial neural networks based chemistry-mixing subgrid model for les. AIAA-2009-241, 47th

AIAA Aerospace Science Meeting, Orlando, FL, 2009.


17 B. Sen, E.R. Hawkes, and S. Menon. Large eddy simulation of extinction reignition with artificial neural networks based

chemical kinetics. Combustion and Flame, 157, 2010.


18 C. C. Rasmussen, J. F. Driscoll, C. D. Carter, M. R. Gruber, J. M. Donbar, and K. Y. Hsu. Blowout limits of supersonic

cavity-stabilized flames. AIAA-2005-3660, 40th AIAA/ASME/SAE/ASEE Joint Propulsion Conference and Exhibit, 2004.
19 M. D. Slessor, M. Zhuang, and P. E. Dimotakis. Turbulent shear-layer mixing: growth-rate compressibility scaling. J.

Fluid Mech., 414:35–45, 2000.


20 E. M. Fernando and S. Menon. Two-dimensional shear-layer entrainment. AIAA, 24:1791–1796, 1993.
21 C. C. Rasmussen, S. K. Dhanuka, and J. F. Driscoll. Visualization of flameholding mechanisms in a supersonic combustor

using plif. Proceedings of the Combustion Institute, 30:2505–2512, 2007.


22 A. Ern and V. Giovangigli. Fast and accurate multicomponent transport property evaluation. J. of Computational

Physics, 120, 1995.


23 F. Genin and S. Menon. Simulation of turbulent mixing behind a strut injector in supersonic flow. AIAA Journal,

48(3):526–539, 2010.
24 B. Sen and S. Menon. Turbulent premixed flame modeling using artificial neural networks based chemical kinetics.

Proceedings of the Combustion Institute, 32, 2009.


25 Kee R. J. Peters, N. The computation of stretched laminar methane airdifussion flames using a reduced four-step

mechanism. Combustion and Flame, 68, 1987.


26 M. Shimada H. Yamashita and T. Takeno. A numerical study on flame stability at the transition point of jet diffusion

flames. Proceedings of the Combustion Institute, 26, 1996.


27 H. Pitsch. Large eddy simulation of turbulent combustion. Annual Review of Fluid Mechanics, 38, 2006.

16 of 16

American Institute of Aeronautics and Astronautics

View publication stats

You might also like