Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

in the reduced mechanism.

For the same reason, the consumption of O in the remaining portion of the combustor is
relatively sharp when compared to detailed mechanism predictions. The consumption of H radical in the separation
region is gradual unlike the detailed mechanism, where it was consumed rapidly by the recombination reactions. Just
downstream of the wall injection location, H radical is consumed faster and afterwards it is gradual. Further, the mass
fraction of all the intermediate species leaving the combustor is higher with reduced mechanism. All the above trends
are expected to have significant impact on the prediction of heat release, specifically in the diverging section.
1.8
Global Reaction
1.7 Reduced Mechanism
1.6 Detailed Mechanism

1.5

T0 / T0,inlet
1.4
1.3
1.2
Downloaded by TEXAS A & M UNIVERSITY on May 9, 2013 | http://arc.aiaa.org | DOI: 10.2514/6.2009-716

1.1
1.0
0.9
-300 -100 100 300 500 700
x (mm)

Figure 4. Variation of dimensionless stagnation temperature with single step,9 detailed mechanism10 and reduced mechanism6 (strut+wall
injection scheme)

Axial variation of dimensionless stagnation temperature along the entire length of the combustor with three different
chemistry models is shown in Fig. 4. In the strut region, fuel injected at lower stagnation temperature mixes with the
flow at higher stagnation temperature (due to heat addition in the separation region ahead of strut injector) resulting
in decrease in stagnation temperature below the inlet value. This drop in stagnation temperature is predicted to be
different with chemistry models. The heat release in the separation region with single step chemistry is the highest and
consequently the drop in stagnation temperature is the lowest. But, the reduced mechanism shows an opposite trend
and so the drop in stagnation temperature is maximum. Further downstream, in the diverging section (where the flow is
kinetically controlled, due to drop in static temperature) the heat release is predicted to be the same with all the chem-
istry model. This in turn highlights the fact that in mixing controlled region the detailed chemistry model is important
to predict the heat release accurately. Heat release with detailed chemistry is higher in the diverging section, which
results in longer separation region (extending upstream of wall injection up to x ≈ 220 mm) as discussed earlier. This
correlates well with the steep decrease in the mass fractions of H, HO2 and H2 O2 due to recombination reactions in
this region in Fig 3. This is accompanied by more heat release in this region. This is in contrast to reduced mechanism
predictions, where the heat release is of lesser magnitude and takes place over a shorter distance. This can be attributed
to higher formation of OH and H radical than that are being consumed, due to the absence of recombination reactions
involving HO2 and H2 O2 , which in turn results in lesser heat release with the reduced mechanism.
Beyond x = 400 mm the rise in stagnation temperature predicted with single step chemistry and reduced mechanism
levels off indicating that the flow is chemically frozen beyond this location. Detailed chemistry calculations, on the
other hand, show the stagnation temperature to increase continuously till the combustor exit, indicating residual heat
release from recombination reactions. The difference in the predictions of stagnation temperature between the chem-
istry models increases continuously from x ≈ 400 mm till the combustor exit. In this region the detailed chemistry
predictions and hence the gradients are higher. Further, the prediction of the stagnation temperature with reduced
mechanism is slightly lower than the single step chemistry prediction. The above discussion clearly underscores the
need for detailed mechanism with recombination reactions for predicting the heat release accurately, which is impor-
tant as it leads to a commensurate augmentation of thrust.
Combustion efficiency is one of the key performance metrics used to evaluate a combustor. Figure 5 shows the compar-
ison of combustion efficiency calculations with reduced mechanism with that of the detailed mechanism10 and single
step chemistry.9 In the multi step chemistry models, since H2 can be formed in the intermediate reactions, combustion
efficiency has to be calculated based on the amount of H2 O formed, rather than the amount of H2 consumed. Thus,

6 of 9

American Institute of Aeronautics and Astronautics Paper AIAA-2009-0716


in the reduced mechanism. For the same reason, the consumption of O in the remaining portion of the combustor is
relatively sharp when compared to detailed mechanism predictions. The consumption of H radical in the separation
region is gradual unlike the detailed mechanism, where it was consumed rapidly by the recombination reactions. Just
downstream of the wall injection location, H radical is consumed faster and afterwards it is gradual. Further, the mass
fraction of all the intermediate species leaving the combustor is higher with reduced mechanism. All the above trends
are expected to have significant impact on the prediction of heat release, specifically in the diverging section.
1.8
Global Reaction
1.7 Reduced Mechanism
1.6 Detailed Mechanism

1.5

T0 / T0,inlet
1.4
1.3
1.2
Downloaded by TEXAS A & M UNIVERSITY on May 9, 2013 | http://arc.aiaa.org | DOI: 10.2514/6.2009-716

1.1
1.0
0.9
-300 -100 100 300 500 700
x (mm)

Figure 4. Variation of dimensionless stagnation temperature with single step,9 detailed mechanism10 and reduced mechanism6 (strut+wall
injection scheme)

Axial variation of dimensionless stagnation temperature along the entire length of the combustor with three different
chemistry models is shown in Fig. 4. In the strut region, fuel injected at lower stagnation temperature mixes with the
flow at higher stagnation temperature (due to heat addition in the separation region ahead of strut injector) resulting
in decrease in stagnation temperature below the inlet value. This drop in stagnation temperature is predicted to be
different with chemistry models. The heat release in the separation region with single step chemistry is the highest and
consequently the drop in stagnation temperature is the lowest. But, the reduced mechanism shows an opposite trend
and so the drop in stagnation temperature is maximum. Further downstream, in the diverging section (where the flow is
kinetically controlled, due to drop in static temperature) the heat release is predicted to be the same with all the chem-
istry model. This in turn highlights the fact that in mixing controlled region the detailed chemistry model is important
to predict the heat release accurately. Heat release with detailed chemistry is higher in the diverging section, which
results in longer separation region (extending upstream of wall injection up to x ≈ 220 mm) as discussed earlier. This
correlates well with the steep decrease in the mass fractions of H, HO2 and H2 O2 due to recombination reactions in
this region in Fig 3. This is accompanied by more heat release in this region. This is in contrast to reduced mechanism
predictions, where the heat release is of lesser magnitude and takes place over a shorter distance. This can be attributed
to higher formation of OH and H radical than that are being consumed, due to the absence of recombination reactions
involving HO2 and H2 O2 , which in turn results in lesser heat release with the reduced mechanism.
Beyond x = 400 mm the rise in stagnation temperature predicted with single step chemistry and reduced mechanism
levels off indicating that the flow is chemically frozen beyond this location. Detailed chemistry calculations, on the
other hand, show the stagnation temperature to increase continuously till the combustor exit, indicating residual heat
release from recombination reactions. The difference in the predictions of stagnation temperature between the chem-
istry models increases continuously from x ≈ 400 mm till the combustor exit. In this region the detailed chemistry
predictions and hence the gradients are higher. Further, the prediction of the stagnation temperature with reduced
mechanism is slightly lower than the single step chemistry prediction. The above discussion clearly underscores the
need for detailed mechanism with recombination reactions for predicting the heat release accurately, which is impor-
tant as it leads to a commensurate augmentation of thrust.
Combustion efficiency is one of the key performance metrics used to evaluate a combustor. Figure 5 shows the compar-
ison of combustion efficiency calculations with reduced mechanism with that of the detailed mechanism10 and single
step chemistry.9 In the multi step chemistry models, since H2 can be formed in the intermediate reactions, combustion
efficiency has to be calculated based on the amount of H2 O formed, rather than the amount of H2 consumed. Thus,

6 of 9

American Institute of Aeronautics and Astronautics Paper AIAA-2009-0716


47th AIAA Aerospace Sciences Meeting Including The New Horizons Forum and Aerospace Exposition AIAA 2009-716
5 - 8 January 2009, Orlando, Florida

47th AIAA Aerospace Sciences Meeting, Orlando, Florida, January 2009.

A Comparison of Numerical Predictions of Supersonic


Combustion of Hydrogen using different Chemistry models in a
Model Combustor
K. Kumaran∗ and V. Babu†
Department of Mechanical Engineering
Indian Institute of Technology
Madras, INDIA 600 036.
Downloaded by TEXAS A & M UNIVERSITY on May 9, 2013 | http://arc.aiaa.org | DOI: 10.2514/6.2009-716

The effect of reduced chemistry model on the numerical predictions of supersonic combustion of hydrogen
in a model combustor is investigated. To this end, 3D, compressible, turbulent, reacting flow calculation with
a reduced chemistry model (with 7 reactions and 7 species) has been carried out. The results are compared
with earlier results obtained using the single step chemistry and detailed chemistry model. Staged injection is
considered to evaluate the capability of the reduced chemistry model. Predictions of mass fractions of major
species, minor species, dimensionless stagnation temperature and combustion efficiency along the combustor
length are presented and discussed. Further, dimensionless wall static pressure is compared with experimen-
tal data reported in the literature. Results show that the reduced chemistry predicts the overall combustion
parameters reasonably well. The heat release is predicted to be lower than that of detailed chemistry model.
Also, the flow separation due to heat release is predicted to be shorter with the reduced mechanism. In this
work, the difference in predictions of supersonic combustion of hydrogen with different chemistry models are
discussed.

I. Introduction
Renewed interest in hypersonic flights has motivated research effort towards developing a supersonic combustion
ramjet (scramjet) engine. Supersonic combustion is the key enabling technology for sustained hypersonic flights. In
scramjet engines the combustor length will typically be of the order of 1m and the residence time of the mixture will
be of the order of milliseconds. This in turn demands effective injection strategies for rapid mixing of fuel with the
mainstream from the combustor design aspect and high diffusivity and short ignition delay from the fuel side. From
the latter perspective, hydrogen is preferred over hydrocarbon fuels due to its high mass diffusivity, wide flammability
limits, minimum ignition energy and low molecular weight. Many investigations (both experimental as well as numer-
ical) have been reported in the literature on the supersonic combustion of hydrogen using different injection strategies.
Modelling of such high speed reacting flows continues to be challenging. Consequently, several chemical kinetics
have been proposed for the combustion of hydrogen. A brief review of the investigations that have been carried out
using different chemistry models for combustion of hydrogen at high speeds is presented next.
Jachimowski1 developed a detailed mechanism for H2 -air reaction and validated with shock tube and laminar flame
studies reported in the literature. Later this mechanism was used to study the effect of chemical kinetics on supersonic
combustion at high Mach numbers. It was reported that at flight speeds (of M = 8), the ignition delay was primarily
controlled by chain-branching and propagation reactions. Further, the influence of recombination reactions and the
inclusion of HO2 chemistry in the mechanism were shown to be crucial in the combustion calculations. In addition,
the prediction of heat release and in turn the temperature in the combustor were shown to be significantly affected by
the rate of formation and consumption of HO2 reactions. Sung et al.2 have shown experimentally that the variation
of ignitability of the H2 − O2 mixture with total pressure at constant total temperature was similar to the classical
homogenous explosion limits of hydrogen-oxygen system. This similarity highlighted the competition between chain-
branching and chain-termination reactions and in turn the inadequacy of the global reaction approximation to predict
∗ Ph.D
Scholar
† Associate
Professor; Corresponding author. E-mail: vbabu@iitm.ac.in
Copyright °
c 2009 by the American Institute of Aeronautics and Astronautics, Inc. All rights reserved.

1 of 9

American
Copyright © 2009 by the American Institute of Aeronautics Institute of Aeronautics
and Astronautics, Inc. All rightsand Astronautics Paper AIAA-2009-0716
reserved.
ignition delay/heat release pattern.
Montgomery et al.3 investigated the effect of reduced mechanism on the predictions of combustion of hydrogen (in-
jected at sonic flow condition) into a Mach 1.9 coaxial air jet. The reduced mechanism predictions were compared
with that of the detailed mechanism and Jachimowski’s mechanism1 without HO2 and H2 O2 reactions. The reduced
mechanism showed good agreement with the detailed mechanism predictions and experimental data. However, the
mechanism without the HO2 and H2 O2 species was shown to exhibit significant difference in the predictions of
species mass fractions. Davidenko et al.4 investigated the effect of modelling parameters on supersonic combustion
of hydrogen in a duct. A reduced mechanism was developed and the predicted results were compared with that of
the predictions with two reduced mechanisms5, 6 and a comprehensive mechanism.7 The predictions with the reduced
mechanism were shown to be comparable with that of the detailed mechanism. However, it was also reported that the
reduced mechanisms might fail in predictions of ignition delay/heat release rate when the temperature of the mixing
layer is not high enough. Also, several other parameters, like, wall static pressure, static temperature, OH species
concentration and axial velocity predictions compared reasonably well with that of the experimental data. Conaire et
al.8 developed a detailed kinetic mechanism to simulate the combustion of H2 − O2 mixtures over a range of pres-
sure, temperature and equivalence ratios. Prediction of ignition delay, flame speed and concentration profiles were
compared with that of the experimental data. Based on a detailed sensitivity study, the three body (HO2 formation
Downloaded by TEXAS A & M UNIVERSITY on May 9, 2013 | http://arc.aiaa.org | DOI: 10.2514/6.2009-716

from H and O2 ) reaction was reported to be highly sensitive among the reactions. Also, it was reported that reactions
involving H2 O2 and HO2 must be taken into account to successfully simulate a wide range of conditions.
Recently, Rajasekaran and Babu9 numerically investigated the supersonic combustion of hydrogen in a model com-
bustor using single step chemistry model. The discrepancy between the numerically predicted wall static pressure and
the experimental data was attributed to over-prediction of mixing by the one equation turbulence model (SA model)
and the inadequacy of the single step chemistry model. The latter issue has been addressed by Kumaran and Babu.10
The effect of detailed chemistry11 on the numerical predictions of supersonic combustion of hydrogen in the same
model combustor have been carried out. The heat release pattern was shown to be significantly different between
single step and detailed chemistry models inside the combustor, wherein the latter model showed higher heat release.
Consequently, higher heat release with detailed chemistry led to a commensurate augmentation of thrust and higher
combustion efficiency in all the injection schemes studied. Also, the residual heat release through recombination re-
actions was found to be significant with detailed chemistry model. Nevertheless, the computational time with detailed
chemistry was considerably higher than that of the single step chemistry model.
All the above studies clearly underscore the fact that the overall combustion parameters of supersonic combustion of
hydrogen can be predicted reasonably well with global reaction approximation. But, in order to improve the combus-
tor design, the intricate details of the reacting flow field, such as the intermediate species distribution, formation and
consumption of these species and the correlation with heat release pattern, pressure rise due to combustion and thrust
profile are needed which are all difficult, if not impossible to measure (except pressure rise). This in turn highlights
the importance of multi step chemical kinetics to accurately predict the ignition delay and heat release in supersonic
combustion. Numerical investigations with detailed/reduced chemistry model can provide insight into these features
of combustion phenomena at high speeds. However, the computational time required for detailed chemistry is too high
to be used for quick evaluation of different combustor designs. This in turn demands an optimized/reduced chemical
kinetics to predict the overall features as well as the intricate details to greater accuracy. The present work is aimed
at studying the effect of reduced chemistry model6 on the predictions of supersonic combustion of hydrogen and to
compare the predictions with that of the predictions with global reaction9 and detailed mechanism.10 Also, a re-
duced mechanism without the H2 O2 and HO2 reactions have been considered to study the influence of recombination
reactions on the predictions.

II. Computational Methodology


The model combustor considered in the present work is the same as the one studied experimentally by Tomioka et
al.12 and numerically by Rajasekaran and Babu9 and Kumaran and Babu.10 Complete geometric detail of the model
combustor is given elsewhere.10, 12 A sectional view of the combustor is shown in Fig. 1. Vitiated air (having a mass
fraction of O2 , H2 O and N2 as 0.198, 0.139, and 0.663 respectively) enters the combustor at a Mach number of 2.5
and a stagnation temperature of 1500 K and stagnation pressure of 1 MPa. Hydrogen fuel at 300 K is injected in
transverse direction (in y-direction) into the supersonic vitiated airstream from the strut and wall injection ports with
an equivalence ratio of 1 (0.4 from strut and 0.6 from wall). There are a total of 10 injectors on the strut (5 each on
top and bottom side of the strut straight region, B-C in Fig. 1) and 8 injectors on the diverging wall (4 each on top and
bottom wall). The injectors are shown by vertical lines in Fig. 1. In this work, all the calculations have been carried

2 of 9

American Institute of Aeronautics and Astronautics Paper AIAA-2009-0716


out on a quarter geometry, from symmetry consideration and hence, fuel is injected only from two and a half ports
on the strut and from two ports on the top wall. In the simulations, three dimensional, compressible and turbulent
Favre-averaged Navier Stokes equations are solved along with species conservation equation.
70
60 H

50
y (mm) 40
30 D F G
20 E
10 B C
A
0
-10
-300 -200 -100 0 100 200 300 400 500 600 700
x (mm)
Downloaded by TEXAS A & M UNIVERSITY on May 9, 2013 | http://arc.aiaa.org | DOI: 10.2514/6.2009-716

Figure 1. Schematic of the model combustor12

In the present work, calculations have been carried out using the one equation, Spalart-Allmaras (SA) model,13 which
is a relatively simple and well established model for aerospace applications, specifically for wall bounded flows. The
SA model with default model constants (Cb1 = 0.1355, Cb2 = 0.622, Cv1 = 7.1, Cw2 = 0.3 Cw3 = 2.0 and κ =
0.4187) has been used here. The turbulent Schmidt number and Prandtl number are taken to be equal to 0.7 and 0.667
respectively. Sensitivity of the numerical predictions to turbulent Schmidt number for ethylene14 and hydrogen15 fuel
have been studied earlier. Laminar finite rate with reduced chemistry model6 has been used in the present work.
The reduced mechanism along with the Arrhenius model constants and the third body efficiency for the termolecular
reactions are given in Table. 1. The reduced mechanism consists of 7 reversible reactions and 7 species (H2 O, O2 ,
H2 , O, H, OH, and N2 ). Effect of the turbulent fluctuations on the rate of reaction has been neglected, which is
largely insignificant for this type of flow conditions.16 Stiff chemistry calculation with default constants has been
enabled in all the calculations. Mass diffusivity of the mixture is modelled using kinetic theory with default values for
Lennard-Jones characteristic length and energy parameter for the individual species. Viscosity and Cp of the mixture
have been evaluated using mass-weighted-mixing law. For the individual species in the mixture these properties have
been evaluated using Sutherland’s law and fifth order polynomials in temperature respectively.

A. Boundary conditions
At the combustor inlet where the flow is supersonic, static and stagnation pressure, stagnation temperature and species
mass fractions are specified. In addition, turbulence intensity (10%) and hydraulic diameter at the inlet have been
specified. At the hydrogen inlet, where the flow is sonic, mass flow rate of hydrogen, static pressure, total temperature
and species mass fraction are specified. The inlet static and stagnation pressures of hydrogen fuel injection port are
adjusted so as to achieve sonic injection with the desired fuel mass flow rate. At the combustor exit where the flow is
supersonic, all the flow variables including pressure are determined from the interior of the domain by extrapolation.
All the walls are considered stationary, adiabatic and no-slip boundary conditions along with standard wall functions
have been used.

B. Grid Independence Study


All the numerical simulations in the present work are being carried out on a well refined mesh used by Rajasekaran
and Babu.9 The final mesh used in their work was obtained after several grid adaption based on gradients of static
pressure and species concentration to resolve the flow features and combustion phenomena accurately. After progres-
sive refinement, a final grid with 751690 cells was used in their work. In this investigation, the maximum value of
wall y + on this grid is 140, while the area-weighted average is 68. Except for some isolated spots in the vicinity of
the injection ports, wall y + is generally between 40 and 100. The calculations are carried out till the global (overall)
mass, momentum and energy balance is within acceptable limits. For the calculations reported here, the difference in
overall mass flow rate between the inlet(s) and outlet is less than 1% of the injected fuel mass flow rate (fuel mass
flow rate rather than the inlet mass flow rate is used since the former is two orders of magnitude less than the latter).
Furthermore, the difference in mass flow rate of H and O atoms between the inlet(s) and the outlet is less than 2% and

3 of 9

American Institute of Aeronautics and Astronautics Paper AIAA-2009-0716


fuel leaving the combustor is almost 35% higher than that of detailed chemistry model. This correlates well with the
mass fraction profiles of H2 O. However, with single step the peak value predicted is almost same as that of detailed
mechanism. This is due to the global reaction approximation (where the combustion of hydrogen produces only water
vapor). Also, this is assisted by the over prediction of static temperature by the SA model,17 which in turn results in
faster consumption of fuel in the strut region.
In the strut region, the reduced mechanism predicts higher H2 O mass fraction than that of the detailed mechanism.
This is attributed to more reaction path ways in detailed mechanism, where the chain branching and propagation re-
actions are dominant due to heat release in the recirculation region (owing to sudden expansion due to a small step
on the strut, shown as B-C in Fig.1) upstream of the strut injection. This results in formation of more intermediate
radicals than the end product with detailed mechanism. However, this trend changes as the flow approaches the wall
injector. The heat addition in an accelerating flow in the diverging section results in flow separation ahead of the wall
injector. This results in availability of H2 and in turn combustion in the separation region ahead of wall injection,
due to which, increase in H2 O and a commensurate decrease in O2 mass fraction profiles are seen in the diverging
section. Near the wall injection location, the length of separation region predicted with detailed mechanism is longer
than that of the reduced mechanism. This is due to higher heat release with detailed mechanism, which in turn results
in longer separation region upstream of the wall injection. This is corroborated well with non-dimensionalized heat
Downloaded by TEXAS A & M UNIVERSITY on May 9, 2013 | http://arc.aiaa.org | DOI: 10.2514/6.2009-716

release plot. Consequently, the mass fraction of H2 O with detailed mechanism starts increasing well ahead of wall in-
jection location compared to reduced mechanism. Downstream of the wall injection location, difference in H2 O mass
fraction between the detailed and reduced mechanism increases till the combustor exit. This is due to higher residual
heat release through recombination reactions with detailed mechanism. Also, the location of H2 O peak predicted by
reduced mechanism lies in between the single step and the detailed mechanism predictions, where the peak location
is predicted further downstream. This in turn reveals that the ignition distance predicted with detailed mechanism is
higher compared to reduced mechanism. The mass fraction profiles of O2 are consistent and in opposite trend to that
of the H2 O mass fraction profile.
10-2 10-2
OH OH

O
10-3 O 10-3
Mass fraction

H H

10-4 10-4

HO2,H2O2
10-5 10-5
-300 -100 100 300 500 700 -300 -100 100 300 500 700
x (mm) x (mm)

Figure 3. Comparison of variation of intermediate species mass fraction along the combustor predicted with detailed mechanism10 (left)
and reduced mechanism6 (right)

Figure 3 shows the variation of minor species along the combustor predicted with detailed mechanism (left) and re-
duced mechanism (right). The peak values of all the intermediate species predicted with reduced mechanism are
higher than that of the detailed mechanism. This is due to less number of intermediate species and chain branching
and propagating reactions in the reduced mechanism. Also, the peak locations are predicted to be slightly upstream
with reduced mechanism, indicating the difference in prediction of ignition distance between these mechanisms. In
the diverging section up to the separation region ahead of the wall injection location, the mass fraction profiles of
intermediate species with reduced mechanism are very similar to that of the detailed mechanism. In the separation
region, the OH mass fraction increases steeply over a short distance with the reduced mechanism, which correlates
well with the heat release pattern. Further downstream, it decreases continuously with reduced mechanism in contrast
to detailed mechanism. The plateau region in the O mass fraction profile ahead of the wall injection location is found
to be shorter with reduced mechanism compared to that of detailed mechanism. This is due to higher consumption of
O radical than formation (i.e., the number of O radical consumption and OH radical formation reactions are higher
than that of O radical formation reactions) due to the absence of recombination reactions involving HO2 and H2 O2

5 of 9

American Institute of Aeronautics and Astronautics Paper AIAA-2009-0716


in the reduced mechanism. For the same reason, the consumption of O in the remaining portion of the combustor is
relatively sharp when compared to detailed mechanism predictions. The consumption of H radical in the separation
region is gradual unlike the detailed mechanism, where it was consumed rapidly by the recombination reactions. Just
downstream of the wall injection location, H radical is consumed faster and afterwards it is gradual. Further, the mass
fraction of all the intermediate species leaving the combustor is higher with reduced mechanism. All the above trends
are expected to have significant impact on the prediction of heat release, specifically in the diverging section.
1.8
Global Reaction
1.7 Reduced Mechanism
1.6 Detailed Mechanism

1.5

T0 / T0,inlet
1.4
1.3
1.2
Downloaded by TEXAS A & M UNIVERSITY on May 9, 2013 | http://arc.aiaa.org | DOI: 10.2514/6.2009-716

1.1
1.0
0.9
-300 -100 100 300 500 700
x (mm)

Figure 4. Variation of dimensionless stagnation temperature with single step,9 detailed mechanism10 and reduced mechanism6 (strut+wall
injection scheme)

Axial variation of dimensionless stagnation temperature along the entire length of the combustor with three different
chemistry models is shown in Fig. 4. In the strut region, fuel injected at lower stagnation temperature mixes with the
flow at higher stagnation temperature (due to heat addition in the separation region ahead of strut injector) resulting
in decrease in stagnation temperature below the inlet value. This drop in stagnation temperature is predicted to be
different with chemistry models. The heat release in the separation region with single step chemistry is the highest and
consequently the drop in stagnation temperature is the lowest. But, the reduced mechanism shows an opposite trend
and so the drop in stagnation temperature is maximum. Further downstream, in the diverging section (where the flow is
kinetically controlled, due to drop in static temperature) the heat release is predicted to be the same with all the chem-
istry model. This in turn highlights the fact that in mixing controlled region the detailed chemistry model is important
to predict the heat release accurately. Heat release with detailed chemistry is higher in the diverging section, which
results in longer separation region (extending upstream of wall injection up to x ≈ 220 mm) as discussed earlier. This
correlates well with the steep decrease in the mass fractions of H, HO2 and H2 O2 due to recombination reactions in
this region in Fig 3. This is accompanied by more heat release in this region. This is in contrast to reduced mechanism
predictions, where the heat release is of lesser magnitude and takes place over a shorter distance. This can be attributed
to higher formation of OH and H radical than that are being consumed, due to the absence of recombination reactions
involving HO2 and H2 O2 , which in turn results in lesser heat release with the reduced mechanism.
Beyond x = 400 mm the rise in stagnation temperature predicted with single step chemistry and reduced mechanism
levels off indicating that the flow is chemically frozen beyond this location. Detailed chemistry calculations, on the
other hand, show the stagnation temperature to increase continuously till the combustor exit, indicating residual heat
release from recombination reactions. The difference in the predictions of stagnation temperature between the chem-
istry models increases continuously from x ≈ 400 mm till the combustor exit. In this region the detailed chemistry
predictions and hence the gradients are higher. Further, the prediction of the stagnation temperature with reduced
mechanism is slightly lower than the single step chemistry prediction. The above discussion clearly underscores the
need for detailed mechanism with recombination reactions for predicting the heat release accurately, which is impor-
tant as it leads to a commensurate augmentation of thrust.
Combustion efficiency is one of the key performance metrics used to evaluate a combustor. Figure 5 shows the compar-
ison of combustion efficiency calculations with reduced mechanism with that of the detailed mechanism10 and single
step chemistry.9 In the multi step chemistry models, since H2 can be formed in the intermediate reactions, combustion
efficiency has to be calculated based on the amount of H2 O formed, rather than the amount of H2 consumed. Thus,

6 of 9

American Institute of Aeronautics and Astronautics Paper AIAA-2009-0716


0.4
Experimental Detailed Mechanism
Global Reaction Reduced Mechanism

0.3

P / P0,inlet
0.2

0.1
Downloaded by TEXAS A & M UNIVERSITY on May 9, 2013 | http://arc.aiaa.org | DOI: 10.2514/6.2009-716

0
-300 -100 100 300 500 700
x (mm)

Figure 6. Comparison of wall static pressure predicted with single step,9 detailed mechanism10 and reduced mechanism6 with experimental
data

IV. Summary and Conclusions


Supersonic combustion of hydrogen with reduced mechanism in a model combustor for the staged injection scheme
has been numerically investigated. Results obtained with reduced mechanism is compared with that of the single step
and detailed chemistry predictions. In the strut region, the peak locations are predicted to be slightly upstream with
reduced mechanism compared to detailed chemistry model, indicating the difference in ignition distance predictions.
Also, the heat release predicted with reduced mechanism is lower than that of the detailed mechanism. This in turn
predicts shorter separation region in the diverging section. In mixing controlled regions, the heat release prediction
with detailed mechanism is shown to be significantly different from that of the reduced mechanism. This in turn
reveals the importance of recombination reactions to accurately predict the heat release in mixing controlled regions
inside the combustor. Nevertheless, the overall combustion parameters, such as thrust calculations, wall static pressure
and the combustion efficiency can be predicted to an acceptable limit with reduced mechanism with considerably less
computational time.

References
1 Jachimowski, C. J., An Analytical Study of the Hydrogen-Air Reaction Mechanism With Application to Scramjet Combustion, NASA-TP-

2791, Feb. 1988.


2 Sung, C. J., Li, J. G., Yu, G., and Law, C. K., Chemical Kinectics and Self-Ignition in a Model Supersonic Hydrogen-Air Combustor, AIAA

Journal, 37, (2), pp. 208-214, 1999.


3 Montgomery, C. J., Zhao, W., Adams, B. R., Eklund, D. R., and Chen, J.-Y., Supersonic combustion simulations using reduced chemical

kinetic mechanisms and ISAT, AIAA-2003-3547.


4 Davidenko, D., Gokalp, I., Dufour, E., and Magre, P., Numerical Simulation of Hydrogen Supersonic Combustion and Validation of Com-

putational Approach, AIAA-2003-7033.


5 Eklund, D. R., Drummond, J. P., and Hassan, H. A., Calculation of Supersonic Turbulent Reacting Coaxial Jets, AIAA Journal, 28, (9), pp.

1633-1641, 1990.
6 Baurle, R. A., and Girimaji, S. S., Assumed PDF turbulence-chemistry closure with temperature-composition correlations, Combustion and

Flame, 134, pp. 131-148, 2003.


7 Mueller, M. A., Kim, T. J., Yetter, R. A., and Dryer, F. L., Flow Reactor Studies and Kinetic Modeling of the H /O Reaction, International
2 2
Journal of Chemical Kinetics, 31, pp. 113-125, 1999.
8 O’Conaire, M., 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, 36, pp. 603-622, 2004.


9 Rajasekaran, A., and Babu, V., Numerical Simulation of Three-Dimensional Reacting Flow in a Model Supersonic Combustor, Journal of

Propulsion and Power, 22, (4), pp. 820-827, 2006.

8 of 9

American Institute of Aeronautics and Astronautics Paper AIAA-2009-0716


10 Kumaran, K., and Babu, V., Investigation of the Effect of Chemistry Models on the Numerical Predictions of the Supersonic Combustion of

Hydrogen, Combustion and Flame, (under review).


11 Stahl, G., and Warnatz, J., Numerical Investigation of Time-Dependent Properties and Extinction of Strained Methane-and Propane-Air

Flamelets, Combustion and Flame, 85, pp. 285-299, 1991.


12 Tomioka, S., Murakami, A., Kudo, K., and Mitani, T., Combustion tests of a staged supersonic combustor with a strut, Journal of Propulsion

and Power, 17, (2), pp. 293-300, 2001.


13 Spalart, P. R., and Allmaras, S. R., A One-Equation Turbulence model for Aerodynamic flows, AIAA-92-0439, Jan 1992.
14 Baurle, R. A., and Eklund, D. R., Analysis of Dual-Mode Hydrocarbon Scramjet Operation at Mach 4-6.5, Journal of Propulsion and Power,

18, (5), pp. 990-1002, 2002.


15 Rajasekaran, A., and Babu, V., On the Effect of Schmidt and Prandtl Numbers in the Numerical Predictions of Supersonic Combustion,

AIAA-2006-5037.
16 Baurle, R. A., Modeling of High Speed Reacting Flows: Established Practices and Future Challenges, AIAA-2004-267.
17 Kumaran, K., Rajasekaran, A., and Babu, V., A Comparison of Numerical Predictions of the Supersonic Combustion of Hydrogen using

S-A and SST K − ω Models, Progress in Computational Fluid Dynamics (under review).

Table 1. Reduced H2 − O2 reaction mechanism6


Downloaded by TEXAS A & M UNIVERSITY on May 9, 2013 | http://arc.aiaa.org | DOI: 10.2514/6.2009-716

S.No. Reactions Pre- Activation Temperature


Exponential Energy Exponent
factor (J/kmol) B
(cm/mol/s)
R(1) H2 + O2 ) * OH + OH 1.70 × 1013 2.0083 × 1008 0.00
R(2) H + O2 * ) OH + O 1.20 × 1017 6.9036 × 1007 -0.91
R(3) OH + H2 * ) H2 O + H 2.20 × 1013 2.1548 × 1007 0.00
R(4) O + H2 * ) OH + H 5.06 × 1004 2.6317 × 1007 2.67
R(5) OH + OH * ) H2 O + O 6.30 × 1012 4560560 0.00
R(6) H + OH + M * ) H2 O + M 2.21 × 1022 0.00 -2.00
*
R(7) H + H + M ) H2 + M 7.30 × 10 17
0.00 -1.00
Third-Body Efficiencies for termolecular reactions
2.50 for M = H2 and 16.0 for M = H2 O, 1.0 for all other species

9 of 9

American Institute of Aeronautics and Astronautics Paper AIAA-2009-0716

You might also like