+ All Categories
Home > Documents > Supersonic Combustion Models Application for Scramjet Engines

Supersonic Combustion Models Application for Scramjet Engines

Date post: 03-Dec-2023
Category:
Upload: cir
View: 4 times
Download: 0 times
Share this document with a friend
26
American Institute of Aeronautics and Astronautics 1 Supersonic Combustion Models Application for Scramjet Engines Francesco Battista * , Luigi Cutrone , Stefano Amabile and Giuliano Ranuzzi § CIRA Italian Aerospace Research Centre, Capua, 81043, Italy This paper presents the application of the CIRA 3D code C3NS to the simulation of supersonic combustion problems typical of scramjet engines. After a brief code description, a trade-off analysis, carried out in order to find a reduced kinetic mechanism for Air-H 2 combustion, is presented. The selected mechanism has to be computationally not heavy but, at the same time, accurate in the prediction of ignition delay time and adiabatic flame temperature. Furthermore, 2D and 3D flow problems reproducing supersonic mixing and combustion processes have been investigated with the purpose of obtaining a verification of the implemented models and a validation by comparing numerical and available experimental data. In particular, three tests of growing complexity, representative of typical supersonic combustion configuration have been selected. The first one is a typical validation tests of mixing and combustion in a parallel-injection configuration (TC-1). The last two tests are typical scramjet engine applications: the TC-2 is a test case representative of a scramjet combustor and TC-3 is a scramjet wind tunnel model in fuel off and fuel on configuration. The results show a good agreement with experimental data and are encouraging in order to employ the code in more complex applications including the evaluation of propulsion systems performances. Nomenclature Acronyms C3NS Combustion 3D Navier Stokes CFD Computational Fluid Dynamics ER Equivalence Ratio ENO Essentially non oscillatory FDS Flux Difference Splitting HC Hydrocarbon ID Identification LES Large Eddy Simulation NP Not present RANS Reynolds Averaged Navier Stokes Latin M Mach number T Temperature T rh Thrust ρ Density p Pressure m Mass u Velocity Subscript st Stoichiometric ox Oxidizer inf Free-stream * Research engineer, Aerospace Propulsion and Reacting Flow Dept., AIAA member, [email protected] , ph. +390823623378. † Ph.D. Research engineer, Aerospace Propulsion and Reacting Flow Dept., AIAA member, [email protected] , ph. +390823623108. Junior Research engineer, Aerospace Propulsion and Reacting Flow Dept., AIAA member, [email protected] , ph. +390823623844. § Ph.D, Research engineer, Aerospace Propulsion and Reacting Flow Dept, [email protected] . 16th AIAA/DLR/DGLR International Space Planes and Hypersonic Systems and Technologies Conferenc AIAA 2009-7207 Copyright © 2009 by F.Battista. Published by the American Institute of Aeronautics and Astronautics, Inc., with permission. Downloaded by ITALIAN CENTER FOR on March 12, 2014 | http://arc.aiaa.org | DOI: 10.2514/6.2009-7207
Transcript

American Institute of Aeronautics and Astronautics

1

Supersonic Combustion Models Application for Scramjet

Engines

Francesco Battista

*, Luigi Cutrone

†, Stefano Amabile

‡ and Giuliano Ranuzzi

§

CIRA Italian Aerospace Research Centre, Capua, 81043, Italy

This paper presents the application of the CIRA 3D code C3NS to the simulation of supersonic

combustion problems typical of scramjet engines. After a brief code description, a trade-off analysis,

carried out in order to find a reduced kinetic mechanism for Air-H2 combustion, is presented. The

selected mechanism has to be computationally not heavy but, at the same time, accurate in the

prediction of ignition delay time and adiabatic flame temperature. Furthermore, 2D and 3D flow

problems reproducing supersonic mixing and combustion processes have been investigated with the

purpose of obtaining a verification of the implemented models and a validation by comparing numerical

and available experimental data. In particular, three tests of growing complexity, representative of

typical supersonic combustion configuration have been selected. The first one is a typical validation tests

of mixing and combustion in a parallel-injection configuration (TC-1). The last two tests are typical

scramjet engine applications: the TC-2 is a test case representative of a scramjet combustor and TC-3 is

a scramjet wind tunnel model in fuel off and fuel on configuration. The results show a good agreement

with experimental data and are encouraging in order to employ the code in more complex applications

including the evaluation of propulsion systems performances.

Nomenclature

Acronyms

C3NS Combustion 3D Navier Stokes

CFD Computational Fluid Dynamics

ER Equivalence Ratio

ENO Essentially non oscillatory

FDS Flux Difference Splitting

HC Hydrocarbon

ID Identification

LES Large Eddy Simulation

NP Not present

RANS Reynolds Averaged Navier Stokes

Latin

M Mach number

T Temperature

Trh Thrust

ρ Density

p Pressure

m Mass

u Velocity

Subscript

st Stoichiometric

ox Oxidizer

inf Free-stream

* Research engineer, Aerospace Propulsion and Reacting Flow Dept., AIAA member, [email protected], ph. +390823623378.

† Ph.D. Research engineer, Aerospace Propulsion and Reacting Flow Dept., AIAA member, [email protected], ph. +390823623108. ‡ Junior Research engineer, Aerospace Propulsion and Reacting Flow Dept., AIAA member, [email protected] , ph. +390823623844.

§ Ph.D, Research engineer, Aerospace Propulsion and Reacting Flow Dept, [email protected].

16th AIAA/DLR/DGLR International Space Planes and Hypersonic Systems and Technologies Conferenc AIAA 2009-7207

Copyright © 2009 by F.Battista. Published by the American Institute of Aeronautics and Astronautics, Inc., with permission.

Dow

nloa

ded

by I

TA

LIA

N C

EN

TE

R F

OR

on

Mar

ch 1

2, 2

014

| http

://ar

c.ai

aa.o

rg |

DO

I: 1

0.25

14/6

.200

9-72

07

American Institute of Aeronautics and Astronautics

2

I. Introduction

Nowadays, the most promising technology for drastical reduction of times to destination for long-distance (i.e.

Brussels to Sidney) civil flights is hypersonic air-breathing propulsion. At hypersonic flight speeds, ranging from Mach

5 to 8, turbofan engines need to be replaced by advanced propulsion concepts, like scramjet engines.

In the design process of Hypersonic Transport Vehicles, the extensive application of CFD tools is essential. The flow

paths envisaged for future scramjet engines will be highly three-dimensional and this requires the application of CFD

codes that are specifically developed for scramjet propulsion, including complex models to accurately describe relevant

physical and chemical effects1.

The aim of this work is to present the application of a 3D numerical solver for the simulation of chemically reactive

flows in typical scramjet engines operating conditions.

The code used in this work is the density-based module of the CIRA code C3NS2. It is able to solve RANS equations

for a generic mixture of gases for steady, subsonic-to-hypersonic flows around 2D, axis-symmetric or 3D geometries.

The correct modeling of the supersonic combustion processes is crucial in the simulation of a scramjet device;

therefore, a trade-off analysis and the selection of proper reduced combustion kinetic mechanisms for supersonic

combustion application is required. In fact, on one hand, due the short resident time of air and propellant inside the

combustion chamber, is necessary to correctly evaluate chemical non equilibrium effects. On the other hand, in order to

reduce CFD computational costs, the selection of a reduced kinetic schemes is highly preferable. Following these

considerations some reduced kinetic schemes for Air-H2 combustion have been analyzed (such as the 8-step model of

Evans and Schexnayder3, the 7-step mechanism of Jachimowski

4 and the single step mechanism of Marinov,

Westbrook and Pitz5) After this analysis, the code has been tested in typical scramjet configurations: some flow

problems have been selected from the relevant literature in order to obtain a verification of the implemented models and

a validation by comparing numerical and available experimental data.

In particular, 2D and 3D flow problems reproducing supersonic mixing and combustion processes have been

investigated:

� TC-1: Parallel Injection (mixing and combustion).

� TC-2: 2D and 3D supersonic combustion in rectangular channel (mixing and combustion).

� TC-3: 3D scramjet wind-tunnel model simulations.

For the last two cases some interesting aspects have been investigated, as the differences between 3D and 2D flow

features and the effects on the computed flowfield of the accuracy of the kinetic scheme in the prediction of the ignition

delay time.

II. C3NS Code Description

The CIRA code C3NS is a structured multiblock finite volume solver that allows the treatment of a wide range of

compressible fluid dynamics problems. The Chemkin II input interface allows to treat different mixture of reacting

gases specifying mixture composition and chemical kinetic scheme. The fluid can be treated as a mixture of perfect

gases in chemical non equilibrium. The thermodynamic model treats mixture of thermally perfect gases, whose specific

heats, enthalpy and entropy are computed by using Gordon McBride polynomial fits9. Species viscosity and thermal

conductivity, are calculated by means of the Eucken law10

, whereas the mixture viscosity and thermal conductivity are

calculated by using the semi-empirical Wilke mixing rules10

. The diffusion coefficients are computed through a sum

rule of the binary diffusivities for each couple of species. In the numerical formulation, the governing equations are

written in integral form, and discretized with a finite volume, cell centred technique. Eulerian fluxes are computed with

a second order Flux Difference Splitting method11

with a ENO reconstruction of the interface values. Viscous fluxes are

computed with a classical centred scheme. Time integration is performed by employing an explicit Euler forward

algorithm coupled with an implicit evaluation of the source terms in time marching approach.

III. Air-H2 Combustion: Kinetic Mechanisms Trade-Off Analysis.

In the study of supersonic combustion problems for advanced propulsion systems, a crucial role is played by

combustion models. The description of combustion processes in a scramjet engine requires finite-rate chemistry

models, because residence times are of the same order of magnitude (typically few milliseconds) as the chemical

characteristic times. Chemical kinetic schemes must be carefully selected, by taking into account the accuracy on one

hand and the computational costs on the other hand. A trade-off analysis of kinetic mechanisms for Air-H2 combustion

Dow

nloa

ded

by I

TA

LIA

N C

EN

TE

R F

OR

on

Mar

ch 1

2, 2

014

| http

://ar

c.ai

aa.o

rg |

DO

I: 1

0.25

14/6

.200

9-72

07

American Institute of Aeronautics and Astronautics

3

has been carried out**

in order to select a reduced kinetic mechanism able to correctly reproduce ignition delay time and

adiabatic flame temperature. The Jachimowski ‘8813 detailed scheme is considered here as reference for the selected

reduced schemes. In Figure 1 a comparison between results obtained by the scheme of Jachimowski and experimental

data in term of ignition delay is reported.

Figure 1 - Validation of Jachimowski ’88 detailed mechanism at p=1 bar13

.

The ranges of pressure (1÷2 bar) and temperature (1100-1300K) selected for the trade-off analysis are typical of

supersonic combustion in scramjet engines. The kinetic mechanisms considered here are the Marinov-Westbrook-Pitz5,

the Jachimowski4 (characterized by 7 steps), and the Evans-Schexnayder

3 schemes. The mixture has been considered

with the ER ranging from 0.5 to 2††

. In Figure 2 the test matrix including all the test conditions considered in the

analysis is graphically shown.

Figure 2 – Test matrix for Air-H2 combustion trade-off analysis.

**

The CIRA SPREAD12 1D code combustion module has been used in this analysis.

††

( )fuel ox

fuel ox st

m mER

m m= .

Dow

nloa

ded

by I

TA

LIA

N C

EN

TE

R F

OR

on

Mar

ch 1

2, 2

014

| http

://ar

c.ai

aa.o

rg |

DO

I: 1

0.25

14/6

.200

9-72

07

American Institute of Aeronautics and Astronautics

4

a) b)

Figure 3 – a) Trade-off analysis results for T=1100K and p=1 bar; b) zoom in timescale.

a) b)

Figure 4 – a) Trade-off analysis results for T=1300K and p=2 bar; b) zoom in timescale.

The results of the analysis, carried out for Air-H2 combustion, are summarized in Figure 3 and Figure 4 . The

Jachimowski (7-steps) mechanism shows the best agreement with the reference scheme, in terms of adiabatic flame

temperature and ignition delay time, in the cases of stoichiometric and fuel-rich mixtures. For low ER cases the

Marinov-Westbrook-Pitz mechanism can be considered a valid alternative to be used in a CFD code, also due to its

simplicity. In the considered conditions the Evans-Schexnayder chemical kinetic scheme overestimates the ignition

delay time and underestimate the adiabatic flame temperature for high initial temperature and pressure, whereas the

combustion does not occur for low initial temperature and pressure.

In this work the effects of the modifications proposed by Craddock6 in order to improve Evans & Schexnayder

mechanism are investigated. To reduce the ignition delay time, the rate coefficients of some reactions has been

changed. In this way, the number of the reactions does not change and the computational efficiency does not decrease

(for figures labels this scheme is called Evans & Schexnayder trebled). The poor modeling of heat distribution with the

reaction model of Evans & Schexnayder is primarily due to the absence of the reactions involving the radical

hydroperoxyl (HO2).

Dow

nloa

ded

by I

TA

LIA

N C

EN

TE

R F

OR

on

Mar

ch 1

2, 2

014

| http

://ar

c.ai

aa.o

rg |

DO

I: 1

0.25

14/6

.200

9-72

07

American Institute of Aeronautics and Astronautics

5

For this reason HO2 radical has been added to the reaction model to improve the heat release modeling capabilities and

the reaction related to this radical have been added to the model (the related scheme is Evans&Schexnayder_8sp). The

subsequent Figure 5 shows the effect of the chemical kinetic scheme modifications.

Figure 5 - Effects on ignition delay time (different time scales).

In the following Figure 6 is shown the comparison of the two scheme results with the other scheme results, where the

enhancement with respect to the original kinetic scheme is evident.

Figure 6 – Comparison with the other kinetic schemes (different time scales).

Dow

nloa

ded

by I

TA

LIA

N C

EN

TE

R F

OR

on

Mar

ch 1

2, 2

014

| http

://ar

c.ai

aa.o

rg |

DO

I: 1

0.25

14/6

.200

9-72

07

American Institute of Aeronautics and Astronautics

6

IV. Supersonic Mixing and Combustion Simulations

In Tab. 1 are listed the simulations carried out with the C3NS code in order to test it in typical flow conditions of a

scramjet engine (other validation tests could be found in Ref.7).

Table 1. Test cases selected

TC Geometry Dim.

1 Parallel injection 2D

2 Supersonic combustion in a rectangular channel 2D – 3D

3 Scramjet tip to tail geometry 2D – 3D

A. TC-1: Parallel Injection The supersonic combustion experiment carried out by Burrows and Kurkov

15-17 has been selected to verify and

validate the code capabilities to simulate a typical Air-H2 parallel injection and combustion in a 2D planar flow.

Figure 7 – TC-1: Geometry and injection configuration

As described in Figure 7, a sonic jet of hydrogen is injected tangentially near the wall into a supersonic vitiated air-

stream flowing into a duct with diverging walls. The jet slot height is 0.4 cm and wall temperature is kept fixed at

298K. The operating conditions are reported in Table 2.

Table 2. TC-1: flow conditions

Air stream H2 Jet

Mach 2.44 1

Temperature (K) 1270 256

Pressure (MPa) 0.1 0.1

Velocity (m/s) 1764 1216

H2 mass fraction 0 1

O2 mass fraction 0.258 0

N2 mass fraction 0.486 0

H2O mass fraction 0.256 0

Dow

nloa

ded

by I

TA

LIA

N C

EN

TE

R F

OR

on

Mar

ch 1

2, 2

014

| http

://ar

c.ai

aa.o

rg |

DO

I: 1

0.25

14/6

.200

9-72

07

American Institute of Aeronautics and Astronautics

7

The two-dimensional computational domain for the numerical simulation of turbulent mixing is divided in 3 blocks and

57267 cells (Figure 8).

Figure 8 – TC-2: computational domain.

The Marinov-Westbrook-Pitz chemical kinetic mechanism has been selected to simulate the hydrogen oxidation

process. This choice can be justified as follows: in Figure 9 are shown, in the mixing zone, just downstream the H2

injection, the equivalence ratio lines ranging from 0.01 to 2 overlapped to the temperature contour lines. According to

Ref.18 the ignition will take place in fuel lean zones (p=0.08 MPa, T>1000K, ER~0.2) where, as discussed in section

III, the Marinov-Westbrook-Pitz mechanism is the best choice among the analyzed ones.

800K 900K1000K

250K

250K

400K600K

800K 800K900K

1100K

1000K

900K

0.01

0.1

11.5

1.521

x[m]

y[m

]

0.11 0.115 0.12 0.125 0.13

0.08

0.085

0.09

0.095

ER: 0.01 0.1 0.5 1 1.5 2

TC-2 : Equivalence Ratio and Temperature contour lines

Figure 9 – TC-2: temperature (black) and ER contour lines in mixing zone.

The calculated mixture fraction profiles for H2 and H2O at 35.6 cm downstream of the injector (at the exit plane see

Figure 11) are compared with the experimental data, respectively in Figure 10 a) and Figure 10 b).

H2

Air

Dow

nloa

ded

by I

TA

LIA

N C

EN

TE

R F

OR

on

Mar

ch 1

2, 2

014

| http

://ar

c.ai

aa.o

rg |

DO

I: 1

0.25

14/6

.200

9-72

07

American Institute of Aeronautics and Astronautics

8

a) b)

Figure 10 – TC-1: Predicted H2 (a) and H2O (b) mole fraction profiles vs. experimental data at x=35.6 cm.

The obtained results present a quite good agreement with experimental data in terms of H2O mole fraction prediction,

even if the flame position is not correctly predicted. In Ref. 15 it is remarked that the flame position is extremely

sensitive to the injection velocity profile used in the flow solver. In this simulation, before the injection in the

combustion chamber, hydrogen passes through a pre-injection duct that gives to the jet a realistic velocity profile before

mixing with air stream. Thus, the lack of accuracy in flame position prediction is probably due to the inlet values of

turbulence quantities, as also stated in Ref.16.

B. TC-2: Supersonic combustion in rectangular channel Differently from the previous test case, the TC-2 is a demonstrative flow case selected to show the code capabilities to

simulate the mixing and the combustion processes with flow features representative of a typical scramjet combustor.

In particular, 2D and 3D simulations of supersonic combustion in a rectangular channel with transverse hydrogen

injection have been carried out. The 2D and 3D geometries, the computational domains and flow conditions are

reported in Figure 11 and Figure 12. The flow conditions have been selected from Von Lavante et al.19

, and assure that

supersonic combustion takes place. Wall temperature is fixed to 293K.

Figure 11 – TC-2: 2D Geometry, grid and flow conditions

19.

Dow

nloa

ded

by I

TA

LIA

N C

EN

TE

R F

OR

on

Mar

ch 1

2, 2

014

| http

://ar

c.ai

aa.o

rg |

DO

I: 1

0.25

14/6

.200

9-72

07

American Institute of Aeronautics and Astronautics

9

Figure 12 - TC-2: 3D Geometry, topology and flow conditions.

Besides the typical flow features of a transverse injection already studied in TC-1 (such as: Prandtl-Meyer expansion,

Mach disk and shock-wave/boundary-layer interaction), this test case is characterized by wall shock reflections, shock-

shock interactions and non-equilibrium combustion chemistry.

In order to demonstrate the flow-solver capabilities to describe such complex phenomena, the case has been initially

simulated by using a two-dimensional 77000-cell (648 in x and 120 in y direction) computational grid divided in 3

blocks. The 3D grid has been obtained by the extrusion of the 2D grid, and has 121 cells in z-direction for a total of

about 9300000 cells in the finest grid level. Its topology is characterized by 6 computational blocks (Figure 12).

The turbulent mixing process has been evaluated by using a classical two-equation k-ε compressible model20

, while

combustion chemical kinetics has been modeled with the Marinov-Westbrook-Pitz single-step mechanism in the 2D

case, and with both the Jachimowski seven steps mechanism4 and Marinov-Westbrook-Pitz single step mechanism in

the 3D case.

2D and 3D tests have been chosen in order to point out differences between 2D and 3D flowfield. Moreover, different

chemical schemes have been used in order to evaluate the effects, on the computed flowfield, of the accuracy in ignition

delay time prediction.

Our investigations begin with the analysis of the results concerning the 2D computations. Figure 13 and Figure 14

respectively provide mirrored views of the computed Mach number and temperature distributions in the combustor

respectively. The complex shock pattern that takes place downstream the injector holes is clearly visible.

Figure 13 – TC-2: Mach number flow field.

Dow

nloa

ded

by I

TA

LIA

N C

EN

TE

R F

OR

on

Mar

ch 1

2, 2

014

| http

://ar

c.ai

aa.o

rg |

DO

I: 1

0.25

14/6

.200

9-72

07

American Institute of Aeronautics and Astronautics

10

A bow shock forms upstream the injector due to the interaction between the supersonic air stream and the hydrogen jet

and reflects on the symmetry axis. Both the bow shock and the reflected shock interact with the boundary layer

respectively upstream and downstream causing shock-induced separations.

Figure 14 – TC-2: Temperature flow field.

As shown in Figure 14, combustion takes place in the region upstream the injector holes, where the mixing occurs at

high temperature. The predicted flame temperature is about 2000 K and the levels of H2O are fairly comparable with

the ones obtained by Von Lavante19

.

A zoom of the separation region is given in Figure 15 where the two recirculation zones, upstream and downstream the

injector hole, are highlighted.

Figure 15 – TC-2: zoom of the separation region (2D case).

In the separation region ahead of the hydrogen transversal jet (see Figure 16), non-zero water concentration can be

found upstream of the jet. In fact, as also observed in Ref. 19, the H2O is carried upstream of the injection opening by

recirculating fluid in the boundary layer. Downstream of the region close to the hydrogen jet, the flow is essentially

chemically frozen, and the H2O is only convected downstream. Moreover, some of the hydrogen still reacts with the

surrounding air in the free shear layer.

Dow

nloa

ded

by I

TA

LIA

N C

EN

TE

R F

OR

on

Mar

ch 1

2, 2

014

| http

://ar

c.ai

aa.o

rg |

DO

I: 1

0.25

14/6

.200

9-72

07

American Institute of Aeronautics and Astronautics

11

Figure 16 – TC-2: 2D H2O mass fraction distribution.

After the analysis of the 2D results, the 3D results will be presented hereinafter. As highlighted in Figure 12 the 3D

geometry is a thin region with a supersonic stream of air in x-direction and with an orthogonal injection of hydrogen.

Although, the geometry is very close to a 2D one, some significant differences must be remarked.

In Figure 17 and Figure 18, respectively, Mach number and temperature contours are shown; respectively the same

flow structures, present in 2D computations, can be recognized. However, in this case, is clearly noticeable the 3D

feature of the flow field. All the flow characteristics, typical of this kind of problem, can be analyzed (by looking also

to Figure 19). The barrel shock, the oblique bow shock, the free shear layer behind the H2 jet and the separation regions

downstream and upstream of the injector are clearly visible. In particular, Figure 18 and Figure 19 show that the flow

forms a horseshoe vortex around the H2 jet, thus relieving the pressure ahead the injection. This feature explains the fact

that the separation regions near the injector hole are smaller in 3D than in 2D (Figure 15). Finally, near the exit plane,

the shock induced separation is not present, because in the 3D configuration the shock reflection is weaker and moved

downward (compare Figure 13 and Figure 17) and the turbulence levels are stronger with respect to the 2D case (as

observed also in Ref.19).

Figure 17 – TC-2: Mach number contours: Section Z=0.02 and X varying sections.

Dow

nloa

ded

by I

TA

LIA

N C

EN

TE

R F

OR

on

Mar

ch 1

2, 2

014

| http

://ar

c.ai

aa.o

rg |

DO

I: 1

0.25

14/6

.200

9-72

07

American Institute of Aeronautics and Astronautics

12

Figure 18 - TC-2: Temperature contours: Section Z=0.02 and X varying sections.

Figure 19 – TC-2 : Zoom of the injector region: Mach contour and streamlines colored with H2 mass fraction.

Dow

nloa

ded

by I

TA

LIA

N C

EN

TE

R F

OR

on

Mar

ch 1

2, 2

014

| http

://ar

c.ai

aa.o

rg |

DO

I: 1

0.25

14/6

.200

9-72

07

American Institute of Aeronautics and Astronautics

13

In this case the predicted flame temperature is around the 2200 K and the levels of H2O production are comparable with

those in Ref. 19.

Figure 20 – TC-2: 3D H2O mass fraction distribution on Z=0.02 m plane.

In particular the levels of H2O in the z-plane symmetry section (Figure 20) are much lower (about 45 %) with respect

to those predicted the 2D case. Figure 21 explains that the H2O distribution is very different with respect to the 2D

case, because of the 3D nature of the flow structures; the H2O is, in fact, in the 3D case, convected downstream of the

injector by the presence of the horseshoe vortex around the H2 jet. This causes the combustion products to round on the

hydrogen jet, thus determining the distribution shown in Figure 21 characterized by the maximum H2O concentration

across the injection.

Figure 21 – TC-3: Zoom of the injector region: H2O mass fraction‡‡

contours and streamlines colored with

Temperature.

A further aspect has been analyzed for TC-3: the effect of kinetic mechanism on the flow field feature. A correct

prediction of ignition delay time is very important in supersonic combustion, because of the small residence time of the

mixture in the combustor.

‡‡

The value of H2O mass fraction is cut off below 0.006, for better visualization purposes, so some zones are blanked.

Dow

nloa

ded

by I

TA

LIA

N C

EN

TE

R F

OR

on

Mar

ch 1

2, 2

014

| http

://ar

c.ai

aa.o

rg |

DO

I: 1

0.25

14/6

.200

9-72

07

American Institute of Aeronautics and Astronautics

14

As stated in the introduction section of this chapter, both Marinov-Westbrook-Pitz and Jachimowski reduced schemes

have been used to perform this 3D test. In chapter III it has been noticed how Marinov-Westbrook-Pitz kinetic

mechanism underpredicts the ignition delay time and the adiabatic flame temperature (Figure 3-Figure 4) with respect

to the detailed one (Jachimowski ‘88). On the contrary, better results have been obtained with the 7-steps Jachimowski

kinetic scheme. In the following figures is shown how the accuracy in ignition delay time prediction could affect the

temperature distribution, shock patterns and H2O production.

Figure 22 –TC-2 : Z=0.02 section. Temperature contour comparison.

Figure 23 –TC-2 : Z=0.02 section constant X planes section. Mach number profiles comparison.

In Figure 22, the temperature contours obtained by 7-steps Jachimowski and Marinov-Westbrook-Pitz kinetic schemes

are compared. It is evident that the results obtained with the simplified mono-step mechanism over-predict the flame

temperature. In fact, in this case, the very underestimated ignition delay time (see section III) causes a stronger

temperature gradient than the one predicted in the computation with the Jachimowski scheme. In Figure 23 Mach

Dow

nloa

ded

by I

TA

LIA

N C

EN

TE

R F

OR

on

Mar

ch 1

2, 2

014

| http

://ar

c.ai

aa.o

rg |

DO

I: 1

0.25

14/6

.200

9-72

07

American Institute of Aeronautics and Astronautics

15

number distributions along y for the sections x1=0.104 m and x2=0.122 m (highlighted in yellow in Figure 22) are

shown. The aim is to show how the shock patterns vary in dependence on the kinetic schemes. In particular the

computation carried out with Marinov-Westbrook-Pitz predicts a steeper shock wave (see Figure 23), which is mainly

due to the different accuracy prediction of the ignition delay time. Moreover, a comparison of the H2O molar fraction

levels is showed in Figure 24, also in this case the simulation with Marinov-Westbrook-Pitz scheme over predicts the

H2O concentration, coherently with the results of section III.

Figure 24 – TC-2: Z=0.02 section. H2O mass fraction contour comparison.

To complete the analysis of TC-3, some comparisons with the available experimental data have been made. In Figure

25 the comparison of the exit velocity of both the 2D and 3D computations with the experimental data is shown.

Differently from 2D, 3D simulations are in good agreement with the experimental data. These results are coherent with

the ones reported in Ref. 19 obtained by Von Lavante using LES methods.

Figure 25 – TC-2: exit velocity comparison with experimental data.

Dow

nloa

ded

by I

TA

LIA

N C

EN

TE

R F

OR

on

Mar

ch 1

2, 2

014

| http

://ar

c.ai

aa.o

rg |

DO

I: 1

0.25

14/6

.200

9-72

07

American Institute of Aeronautics and Astronautics

16

Furthermore, a qualitative comparison of the numerical density gradient compared with the Schlieren experimental

picture is presented. It can be noticed (Figure 26) that the predicted flow structures are similar to those visible in the

experimental flow visualization§§

. This is verified mainly near the injection region, where grid resolution is high, but is

worse in normal direction because the grid rapidly becomes coarser. It is remarkable that the positions of the reflected

shock waves are quite well predicted. It should be actually noticed that the RANS methodology does not allow to

obtain an accurate vortex structures description.

Figure 26 – TC-2: Density gradient comparison (CFD results VS Experimental Schlieren).

C. TC-3: Tip to Tail Scramjet simulations This section reports the 2D and 3D numerical simulations of the shock-tunnel test, carried out in the free piston T4

facility of the Queensland University, of the scramjet-like configuration shown in Figure 27. The experimental

configuration21

involves injection of hydrogen fuel into the scramjet inlet, followed by mixing, shock-induced ignition,

and combustion. This configuration has been subsequently investigated by CFD analysis in the work of Star22

.

Figure 27 - Photograph of the scramjet model mounted in the test section of the T4 shock tunnel21

.

§§

In the figure all the shock patterns are remarked for both experimental and numerical picture in order to clarify the analysis.

Dow

nloa

ded

by I

TA

LIA

N C

EN

TE

R F

OR

on

Mar

ch 1

2, 2

014

| http

://ar

c.ai

aa.o

rg |

DO

I: 1

0.25

14/6

.200

9-72

07

American Institute of Aeronautics and Astronautics

17

The 3D geometry is reported in Figure 28 where one of the sidewall has been removed for illustration proposal. The

scramjet model has a rectangular cross-section consisting of an inlet section with an integrated fuel delivery system, a

combustion chamber and a simply ramped nozzle section. The total length of the model is 0.625 m, the width is 0.075

m, and the combustor section considered here is of 0.024 m. The two inlet ramps are respectively inclined by 7 deg and

9 deg with respect to the horizontal. The fuel delivery subsystem injects choked gaseous hydrogen at a 45 deg angle

(relative to the surface) through four circular discrete holes located on both upper and lower surfaces of the first inlet

ramp section upstream of the combustor section. The four fuel holes are evenly spaced in the spanwise direction, with

the intent to provide the fuel and air with enough time to mix and properly ignite, of course only when the shock

induced temperature rises sufficiently.

Table 3. TC-4: model geometric specifications.

Figure 28 – Three-Dimensional Scramjet Model and Centerline Cross-Section.

The experimental campaign carried out by Odam e Paul21

considered both fuel-off and fuel-on runs. In this work two

representative points of both these categories of experiment have been considered; fuel-off simulation in order to

analyze the air flow feature inside the scramjet engine model and assesses grid topology and refinement, fuel-on

simulation in order to predict all the processes typical of a scramjet thruster such as compression, induced shock

ignition, supersonic combustion and acceleration. In the following tables are reported the conditions that have been

identified with the ID number of the related experimental run21

.

Dow

nloa

ded

by I

TA

LIA

N C

EN

TE

R F

OR

on

Mar

ch 1

2, 2

014

| http

://ar

c.ai

aa.o

rg |

DO

I: 1

0.25

14/6

.200

9-72

07

American Institute of Aeronautics and Astronautics

18

Table 4. Fuel-off run conditions (RUN#767521

)

Air stream H2 Jet

Mach 6.25 NP

Temperature (K) 487 NP

Pressure (Pa) 10230 NP

Velocity (m/s) 2766 NP

H2 mass fraction 0 NP

O2 mass fraction 0.233 NP

N2 mass fraction 0.766 NP

TWall (K) 300

Table 5. Fuel-on run conditions (RUN#767821

)

Air stream H2 Jet

Mach 6.42 1

Temperature (K) 412 250

Pressure (Pa) 8958 647058

Velocity (m/s) 2612 1201

H2 mass fraction 0 1

O2 mass fraction 0.233 0

N2 mass fraction 0.766 0

TWall (K) 300

Computational grids are referred to the quarter-symmetry scramjet geometry showed in Figure 29, where in both the 2D

and 3D computational grids meshed geometries are presented.

Figure 29 – TC-3 : Computational domain.

The 3D grid used is characterized by 38 blocks (Figure 30) and has 2000134 cells on the finest grid level. The

minimum normal spacing for both sidewall and combustor wall is about 1.0e-6 m.

Dow

nloa

ded

by I

TA

LIA

N C

EN

TE

R F

OR

on

Mar

ch 1

2, 2

014

| http

://ar

c.ai

aa.o

rg |

DO

I: 1

0.25

14/6

.200

9-72

07

American Institute of Aeronautics and Astronautics

19

a) b***

)

Figure 30 – TC-3: 3D blocks decomposition (a) and grid (b).

The two dimensional grid was extracted from the mid-plane of the quarter-symmetry three dimensional grid with an

irrelevant rearrangement in the fuel injector zone. The cells number of the finest grid level is 10786 and the stretching

at the wall is the same of the 3D grid, of course.

Figure 31 – TC-3: 2D grid.

Fuel-off computations, as previously said, are carried out in order to preliminary identify main air flow features in the

scramjet geometry and, furthermore, to asses grid refinement effects. In the following, 3D simulations results, in terms

of upper wall pressure distributions, are presented and compared with the 2D ones and with experimental data†††

(see

Figure 32) .

***

For visualization purposes wall fitting has been relaxed. †††

The experimental pressure measurements are obtained from pressure taps located on the centreline of the model21,22

.

Dow

nloa

ded

by I

TA

LIA

N C

EN

TE

R F

OR

on

Mar

ch 1

2, 2

014

| http

://ar

c.ai

aa.o

rg |

DO

I: 1

0.25

14/6

.200

9-72

07

American Institute of Aeronautics and Astronautics

20

Figure 32 – TC-3 : Comparison between numerical and experimental static pressure distribution along the wall

(fuel-off case).

The fuel-off solution on the finest grid level has been analyzed in order to have an accurate description of air flow

features inside the scramjet model. The plot of the magnitude pressure gradient is also showed in Figure 32.

In the Figure 32 the key flow features can be distinguished and described as follows.

A. First inlet ramp shock.

B. Second inlet ramp shock.

C. Corner expansion fan arising from the flow near the wall turning to enter the combustor section.

D. Wall interaction of the first ramp shock with the transmitted shock from the opposite wall.

E. Wall interaction of the second inlet ramp shock with the transmitted shock from the opposite wall.

F. Wall interaction of the expansion fan from C transmitted from the opposite wall.

G. Wall interaction with the transmitted combined shock structure of D and E.

H. Wall interaction of the expansion fan from F transmitted from the opposite wall.

I. Wall interaction with the transmitted shock structure of G.

J. Corner expansion fan arising from the flow near the wall turning to enter the nozzle section.

K. Final transmitted shock system.

Dow

nloa

ded

by I

TA

LIA

N C

EN

TE

R F

OR

on

Mar

ch 1

2, 2

014

| http

://ar

c.ai

aa.o

rg |

DO

I: 1

0.25

14/6

.200

9-72

07

American Institute of Aeronautics and Astronautics

21

Moreover, it can be noticed from Figure 32 the good agreement of the 3D simulation results with the experimental

data. It is remarkable that the 3D computation better predicts the position of the corner expansion fan and the final

shock system with respect to the 2D computation‡‡‡

, as expected.

In conclusion, a detailed view of the 3D flow field is shown in the following figure. In particular, Figure 33 shows the

3D temperature distribution. It can be noticed that the temperature peaks are in proximity of the wall shocks

impingements, moreover the angularity of the flow in the corner of the combustion chamber is visible on the

temperature maps. These recirculation zones will be significant for the air-fuel mixing in the fuel-on conditions.

Figure 33 – TC-4: Fuel-off case 3D. Temperature contour slices.

The main focus of the fuel-on test-case is to asses code capability in the simulation of all the structures typical of a

scramjet engine flow, such as: compression, shock induced ignition, supersonic combustion and expansion. In addition

particular performance parameters such as thrust and wall heat flux have been derived.

In particular, Figure 34 and Figure 35 show that the fuel jet initially penetrates through the thin boundary layer, then the

fuel is turned toward the wall downstream of the initial shock impingement point and resides within the boundary layer

before partially combusting midway through the combustion chamber, where is visible the conversion of the O2 in H2O

(Figure 36).

‡‡‡

Even if there is a good agreement between the computed and the experimental data a further grid refinement could improve the shock structures

resolution.

Dow

nloa

ded

by I

TA

LIA

N C

EN

TE

R F

OR

on

Mar

ch 1

2, 2

014

| http

://ar

c.ai

aa.o

rg |

DO

I: 1

0.25

14/6

.200

9-72

07

American Institute of Aeronautics and Astronautics

22

Figure 34 – TC-3: Fuel-on 3D. H2 penetration (quarter symmetry geometry).

Figure 35 – TC-3: Fuel-on 3D, H2O production.

Dow

nloa

ded

by I

TA

LIA

N C

EN

TE

R F

OR

on

Mar

ch 1

2, 2

014

| http

://ar

c.ai

aa.o

rg |

DO

I: 1

0.25

14/6

.200

9-72

07

American Institute of Aeronautics and Astronautics

23

Figure 36 – TC-3 : Fuel-on 3D. H2O production and O2 consumption along the wall on symmetry plane.

In Figure 37, the heat flux distribution on the upper wall of the combustion chamber and nozzle is shown; it is evident

that the heat flux peaks occur in correspondence of the shock reflections at the wall.

Figure 37 – TC-3: Fuel-on 3D case. Upper wall heat flux distribution.

The experimental model, analyzed numerically in this section, was designed in order to study combustion phenomena

in scramjet geometry, so the main intent was not to have a model optimized for thrust. Anyway, some evaluations of

thrust levels have been made21

. The experimental total thrust levels presented21

are in terms of thrust coefficient defined

as follows:

Dow

nloa

ded

by I

TA

LIA

N C

EN

TE

R F

OR

on

Mar

ch 1

2, 2

014

| http

://ar

c.ai

aa.o

rg |

DO

I: 1

0.25

14/6

.200

9-72

07

American Institute of Aeronautics and Astronautics

24

2

inf inf inf

1

2

rhT

TC

u Aρ

= ;

Total thrust§§§

could be obtained integrating CFD results in terms of pressure distributions over the inflow and outflow

surfaces of the scramjet model geometry. The thrust coefficient, derived by means of the CFD computation is

0.13TC = , a value that is comparable with the values reported in Ref.21 ( 0.16TC = ) for the combustion chamber

height and equivalence ratio here considered. Finally the comparison between 3D computations and experimental data

are presented in terms of upper wall pressure distribution (Figure 38).

Figure 38 – TC-3: Comparison between numerical and experimental static pressure distribution along the wall

(fuel-on case)

It can be noticed that, even if there is a good agreement in terms of thrust coefficient between experimental data and

CFD computation, the comparison in terms of pressure distribution show a not so good agreement. Further

investigation are needed in terms of turbulence models, chemical scheme selection and grid refinement.

V. Conclusions and Future Works

The interest in scramjet technology in recent years and the successful test flight of hypersonic propulsion vehicles

attests the technical feasibility of this technology. It is well known that a crucial element in the successful development

of hypersonic propulsion device is an efficient supersonic combustor. The supersonic flow conditions in the combustor

lead to very short residence times for fuel and air, typically of the order of few milliseconds; therefore the fuel has to

mix with the supersonic flow and burn completely within such a short time. Also the ignition and the reaction times

have to be short. Hydrogen is suited to meet these requirements. During the development of a viable supersonic

combustor, extensive experimental investigations of combustor models with different injection and flame stabilizations

techniques are usually carried out. In contrast, fully resolved numerical simulations of the 3D turbulent, reacting flow in

the supersonic combustor are limited. This has provided the motivation of this work. In fact, this paper has described

the activities aimed at the verification and application of the density-based module of the 3D CIRA code C3NS,

developed to properly simulate supersonic combustion phenomena.

§§§

This value does not include the thrust contributed by the momentum addition of injecting the fuel; however, the

theoretical analysis suggest that the fuel injection thrust contributes less than 1% of the total fuel-on thrust.

Dow

nloa

ded

by I

TA

LIA

N C

EN

TE

R F

OR

on

Mar

ch 1

2, 2

014

| http

://ar

c.ai

aa.o

rg |

DO

I: 1

0.25

14/6

.200

9-72

07

American Institute of Aeronautics and Astronautics

25

The main findings of present work can be summarized as follows:

� A trade-off analysis of chemical kinetics mechanisms for Air/H2 combustion has been performed in order to

select proper reduced chemical kinetics schemes to be embedded in C3NS for the study of supersonic

combustion.

� Some relevant flow cases, both 2D and 3D, have been selected from literature in order to test the solver

capabilities in predicting mixing and combustion processes in typical scramjet engine injection configurations:

transverse and parallel injection. The numerical results have been compared with available experimental

measurements. From these comparisons we can conclude that the code is able to predict:

� the complex wave pattern and separation caused by transverse gas injections in a turbulent supersonic

air stream, characterized by the presence of a bow shock upstream the injector, a rapid Prandtl-Mayer

expansion, Mach disk formation and shock wave-boundary layer interactions;

� mixing and combustion in supersonic planar shear layers generated by parallel injection of gaseous

hydrogen in a supersonic air flow;

� mixing and combustion processes in 2D and 3D channels (representative of scramjet combustion

chambers) with hydrogen transverse injection, in presence of multiple shock reflections and shock-

induced separations.

� Simulations of a scramjet wind-tunnel model have been performed in order to verify C3NS solver in cases

including flow features typical of a whole scramjet engines.

The test carried out on the scramjet wind tunnel model has shown that tests with a further grid refinement will be

necessary; furthermore has been highlighted the importance of chemical kinetic scheme selection, so other kinetic

scheme performances will be investigated. Future computational activities will be focused in using C3NS to support the

design of a real model of scramjet engine, in order to evaluate flow features and performance parameters such as thrust

and wall thermal loads.

Acknowledgments

This work was performed within the “Long-Term Advanced Propulsion Concepts and Technologies” (LAPCAT)

project investigating high-speed air-breathing propulsion. LAPCAT, coordinated by ESA-ESTEC, is supported by the

EU within the 6th

Framework Programme Priority 1.4, Aeronautic and Space, Contract no.: AST4-CT-2005 012282.

Further info on LAPCAT can be found on http://www.esa.int/techresources/lapcat.

References

1 LAPCAT Project Description of Work, Contract for Specific Target Research and Innovation Project, Annex I,

April 2006.

2 Battista F., Cutrone L., Ranuzzi G., C3NS Solver for Supersonic Combustion Simulations: Final Report, LAPCAT

6.4.4 Deliverable, April 2008, CIRA-CF-08-0484.

3 Evans J. S. and Schexnayder C. J. J., Influence of Chemical Kinetics and Unmixedness on Burning in Supersonic

Hydrogen Flames, AIAA Journal, Vol. 18, No. 2, 1980, pp. 188-193.

4 Drummond, J. P. and Rogers, R. C., A Numerical Model for Supersonic Reacting Mixing Layers, Computer

Methods in Applied Mechanics and Engineering, Vol.64, 1987, pp. 39-60.

5 Marinov-Westbrook-PitzN.M., Westbrook C.K. and Pitz W.J., Detailed and Global Chemical Kinetics Model for

Hydrogen, Transport phenomena in Combustion, Vol.1, Taylor and Francis, Washington DC, 1996.

Dow

nloa

ded

by I

TA

LIA

N C

EN

TE

R F

OR

on

Mar

ch 1

2, 2

014

| http

://ar

c.ai

aa.o

rg |

DO

I: 1

0.25

14/6

.200

9-72

07

American Institute of Aeronautics and Astronautics

26

6 Christopher S. Craddock B.E., Computational Optimization of Scramjets and Shock Tunnel Nozzle, Ph.D. thesis

1999.

7 Battista F., Cutrone L., Ranuzzi G. and Borreca S., Supersonic Combustion Models Application in Advanced

Propulsion Concepts, 15th AIAA International Space Planes and Hypersonic Systems and Technologies

Conference, Dayton, AIAA-2008-2553.

8 Park C., Review of Chemical-Kinetic Problems of future NASA Missions, I: Earth Entries, Journal of

Thermophysics and Heat Transfer, Vol.7, No. 3, 1993.

9 Gordon, S., McBride, Computer Program for Calculation of Complex Chemical Equilibrium Compositions,

Rocket performance, Incident and Reflected Shocks and Chapman-Jouguet Detonations, NASA Report SP-273,

1971.

10 Hirshfelder, Curtiss, Bird, Molecular Theory of Gases and Liquids, John Wiley and Sons, New York, 1954.

11 Borrelli S., Pandolfi M., An Upwind Formulation for the Numerical Prediction of Non Equilibrium Hypersonic

Flows, 12th International Conference on Numerical Methods in Fluid Dynamics, Oxford, United Kingdom, 1990.

12 Bonifacio S., Borreca S., Ranuzzi G., and Salvatore V., SPREAD: a Scramjet PREliminary Aerothermodynamic

Design Code, 14th AIAA/AHI Space Planes and Hypersonic Systems and Technologies Conference, Canberra,

Australia, November 2006.

13 Jachimowski, C. J., An Analytical Study of the Hydrogen-Air Reaction Mechanism with Application to Scramjet

Combustion, NASA TP-2791, February 1988.

14 Rogers C.R., Schexnayder C.J.Jr. , Chemical Kinetic Analysis of Hydrogen-Air Ignition And Reaction Times,

NASA-TP-1856,1981.

15 Ebrahimi H. B., An Overview of Computational FluidDynamics for Application to Advanced Propulsion Systems,

11th AIAA/AAAF International Space Planes and Hypersonic Systems and Technologies Conference, Orleans,

France, AIAA 2002-5130.

16 Ebrahimi H. B., Validation Database for Propulsion Computational Fluif Dynamics, Journal of Spacecraft and

Rockets, Vol.34, No.5, September-October 1997.

17 Burrow M., and Kurkov A. P., Analytical and Experimental Study of Supersonic Combustion of Hydrogen in a

Vitiated Airstream, NASA-TMV-2828, September 1973.

18 Huber P.W., Schexnayder C. J. J., McClinton C.R., Criteria for Self-Ignition of Supersonic Hydrogen-Air

Mixture, NASA TP 1457, 1979.

19 Von Lavante E., Zeitz D., and Kallenberg M., Numerical Simulation of Supersonic Airflow with Transverse

Hydrogen Injection, AIAA Journal, Vol.17, No.6, 2001.

20 Grasso F., and Falconi D., High-Speed Turbulence Modeling of Shock-Wave/Boundary-Layer Interaction, AIAA

Journal, Vol. 31, No. 7, July 1993.

21 Odam, J., and Paull, A., Internal Combustor Scramjet Pressure Measurements in the T4 Shock Tunnel, 11th

AIAA/AAAF International Space Planes and Hypersonic Systems and Technologies Conference, Orleans, France,

Sept., 2002. AIAA 2002-5244.

22 Star B. J, Edwards J. R. Jr, Smart M.K, Baurle R.A, Numerical Simulation of Scramjet Combustion in a Shock

Tunnel, 43rd Aerospace Sciences Meeting and Exhibit, AIAA 2005-0428,

Dow

nloa

ded

by I

TA

LIA

N C

EN

TE

R F

OR

on

Mar

ch 1

2, 2

014

| http

://ar

c.ai

aa.o

rg |

DO

I: 1

0.25

14/6

.200

9-72

07


Recommended