Vous êtes sur la page 1sur 9

Proceedings

of the
Combustion
Institute
Proceedings of the Combustion Institute 30 (2005) 2929–2937
www.elsevier.com/locate/proci

Thermophysical characteristics of shear-coaxial


LOX–H2 flames at supercritical pressure
Joseph C. Oefelein*
Combustion Research Facility, Sandia National Laboratories, Livermore, CA 94551-9051, USA

Abstract

The thermophysical characteristics of a liquid-oxygen–hydrogen flame at supercritical pressure are


investigated, with emphasis placed on the near-field region just downstream of a shear-coaxial injector ele-
ment. Results are presented from a series of hierarchical high-fidelity simulations for a condition where
oxygen is injected in a cryogenic state, at a subcritical temperature and supercritical pressure, and hydrogen
is injected in a supercritical state. This condition has significant technical relevance in liquid-rocket engines,
but is not well understood and has not been well quantified to date. For this situation, a diffusion domi-
nated mode of combustion occurs in the presence of exceedingly large thermophysical property gradients.
The flame anchors itself in the interfacial region of high shear that exists between the liquid-oxygen core
and the annular hydrogen jet, and intensifies the respective property gradients such that they approach
the behavior of a contact discontinuity. Significant real-gas effects and transport anomalies coexist locally
in colder regions of the flow, with ideal gas and transport characteristics occurring within the flame zone.
The current analysis provides one of the first quantitative characterizations of the flame structure and asso-
ciated property variations in liquid-oxygen–hydrogen flames at supercritical pressure.
 2004 The Combustion Institute. Published by Elsevier Inc. All rights reserved.

Keywords: Turbulent combustion; Cryogenic propellants; Supercritical pressure; Flame structure; Liquid-rocket
propulsion

1. Introduction including the Space Shuttle Main Engine (SSME)


and the Ariane 5 Vulcain engine. The chamber
1.1. Objectives pressure in these engines exceeds 100 atm, which
is significantly higher than the thermodynamic
Use of cryogenic propellants in liquid-rocket critical pressure of both propellants.
engines at chamber pressures that exceed the ther- Although LOX–H2 rocket engines have been
modynamic critical point of the propellants has operated relatively safely for the past several
been well established as a combination that pro- years, the processes that control combustion are
vides high efficiency and performance for a variety still not well understood. This paper provides
of launch vehicle applications. In particular, li- one of the first quantitative characterizations of
quid-oxygen–hydrogen (LOX–H2) engines have the flame structure and associated property varia-
been developed for a variety of launch systems, tions in LOX–H2 flames at supercritical pressure.
Results from a series of hierarchical simulations
are presented for a condition where oxygen is in-
*
Fax: +1 925 294 2595. jected in a cryogenic state, at a subcritical temper-
E-mail address: oefelei@sandia.gov. ature and supercritical pressure, and hydrogen is

1540-7489/$ - see front matter  2004 The Combustion Institute. Published by Elsevier Inc. All rights reserved.
doi:10.1016/j.proci.2004.08.212
2930 J.C. Oefelein / Proceedings of the Combustion Institute 30 (2005) 2929–2937

injected in a supercritical state. Emphasis is placed ters that subsequently break up and vaporize. As
on shear-coaxial injection processes in the labora- the chamber pressure approaches the thermody-
tory-scale uni-element rocket configuration devel- namic critical pressure of the LOX (Fig. 1B), how-
oped at the German Aerospace Center (DLR) in ever, the number of drops present diminishes, and
Lampoldshausen. the injected LOX jet begins to exhibit a pure dif-
fusion mechanism. This mechanism was observed
1.2. Phenomenological trends to dominate at a pressure of 4.5 MPa, which is
slightly below the thermodynamic critical pressure
Shear-coaxial injection processes in liquid- of oxygen, and significantly above that of hydro-
rocket engines exhibit two distinct modes of com- gen. Studies carried out in the ONERA ‘‘MAS-
bustion [1–5]. At subcritical pressures, the liquid COTTE’’ facility [6,7] have produced similar
jets atomize. Dynamic forces and surface tension results. Flame attachment occurs instantaneously
promote the formation of a heterogeneous spray, after ignition in the small but intensive recircula-
and lifted spray flames form in a manner consis- tion zone that forms just downstream of the annu-
tent with the modes of combustion exhibited by lar LOX-post. Gurliat et al. [8] have recently
local drop clusters. When chamber pressures ap- verified this trend over a broad range of
proach or exceed the critical pressure of a partic- conditions.
ular propellant, however, injected liquid jets
undergo a transcritical change of state as interfa-
cial fluid temperatures rise above the critical tem- 2. Approach
perature of the local mixture. For this situation,
diminished inter-molecular forces promote diffu- Results were obtained using the theoretical–
sion dominated mixing processes prior to atom- numerical framework developed by Oefelein
ization. Respective jets vaporize in the presence [9,10]. This framework has been designed for
of exceedingly large thermophysical gradients, application of both the direct numerical simula-
and well-mixed diffusion flames evolve as a conse- tion (DNS) technique, and the large eddy simula-
quence that are anchored by small but intensive tion (LES) technique and solves the fully coupled
recirculation zones just downstream of the LOX- compressible conservation equations of mass,
post. momentum, total-energy, and species in full
Flow visualization studies conducted by Mayer geometries. A full multicomponent formulation
and Tamura [1] have illustrated the trends de- has been employed for the current investigation.
scribed above for the case of a LOX–H2 shear-co- The system is assumed to be compressible, chem-
axial injector element. The two extremes are ically reacting, and composed of N species.
shown in Fig. 1. When LOX is injected at low- The baseline system of equations are cast in
subcritical pressures (Fig. 1A), atomization occurs dimensionless form using a reference length-scale
forming a distinct spray. Ligaments are detached dref, flow speed Uref, and fluid state characterized
from the jet surface forming drops and drop clus- by a reference density qref, sound speed cref, con-
stant pressure specific heat C pref , and dynamic vis-
cosity lref. Using these quantities, a reference
Mach and Reynolds number are defined as
M = Uref/cref and Re = qrefUrefdref/lref. With these
definitions, the instantaneous conservation equa-
tions of mass, momentum, total-energy, and
chemical species can be written in conservative
form as follows:
oq
þ r  ðquÞ ¼ 0; ð1Þ
ot
 
o p
ðquÞ þ r  qu  u þ 2 I ¼ r  s; ð2Þ
ot M
where
 
l 2  
s¼  ðr  uÞI þ ru þ ruT
Re 3
represents the viscous stress tensor;

o
Fig. 1. Reacting shear-coaxial LOX–H2 injector oper- ðqet Þ þ r  ½ðqet þ pÞu
ating at (A) 1.5 MPa (15 atm) and (B) 4.5 MPa (44 atm). ot
 
From Mayer and Tamura [1]. Used with permission. ¼ r  qe þ r  M 2 ðs  uÞ ; ð3Þ
J.C. Oefelein / Proceedings of the Combustion Institute 30 (2005) 2929–2937 2931

where employed was developed by Oefelein [10] and is


designed to account for thermodynamic non-ide-
M 2 X
N
p
et ¼ e þ u  u; e¼ hi Y i  ; alities and transport anomalies over a wide range
2 i¼1
q of pressures and temperatures. Here, the extended
corresponding states model [14,15] is used with a
and
Z Z cubic equation of state to evaluate the p–v–T
p T
behavior of the inherent dense multicomponent
hi ¼ h0fi þ C pi ðT ; pÞ dT dp
p0 T0
mixtures. A summary of the cubic equations of
state and recommended constants is given by Reid
are the total internal energy, internal energy, and et al. [16, Chapter 3]. Experience [17,18] has
partial enthalpy of the ith species, and shown that both the Soave–Redlich–Kwong
o (SRK) and Peng–Robinson (PR) equations, when
ðqY i Þ þ r  ðqY i uÞ used in conjunction with the corresponding states
ot
principle, can give accurate results over the range
¼ r  qi þ x_ i ; i ¼ 1; . . . ; N  1; ð4Þ of pressures, temperatures, and mixture states of
where x_ i represents the instantaneous rate of pro- interest in this study. SRK coefficients are fit to
duction of the ith chemical species due to reac- vapor pressure data, and thus more suitable when
tions. Heat release due to chemical reactions in the reduced temperature is less than one. PR coef-
the total-energy equation is accounted for in the ficients, on the other hand, are more suitable when
description of the partial specific enthalpies, hi, the reduced temperature is greater than one. Here,
by including the enthalpy of formation, h0fi , in the PR equation of state was used exclusively
the definition. since the flow involves heat release.
The dimensionless multicomponent energy and Relevant thermodynamic properties are ob-
mass diffusion fluxes are given by the expressions tained in two steps. First, respective component
properties are combined at a fixed temperature
lC p XN
using the corresponding states methodology to
qe ¼  rT þ hi qi obtain the mixture state at a given reference
PrRe i¼1
pressure. A pressure correction is then applied
XN X N
X j DTi using departure functions of the form given by
þ Ru T ðVi  Vj Þ ð5Þ Reid et al. [16, Chapter 5]. These functions are
i¼1 j¼1
W i Dij
derived from MaxwellÕs relations (see VanWylen
and and Sonntag [19, Chapter 10], for example) and
make full use of the real-mixture p–v–T path
qi ¼ qY i Vi dependencies dictated by the equation of state.
XN Molecular transport properties are evaluated in
Y iY j
¼ DTi rðln T Þ þ q Dij dj ; ð6Þ an analogous manner. The viscosity and thermal
j¼1
X iX j conductivity are obtained using the extended
corresponding states methodologies developed
where Ru ð¼ Ru =C p ref Þ is the dimensionless form by Ely and Hanley [20,21]. Mass and thermal
of the universal gas constant, Vi is the dimension- diffusion coefficients are obtained using the
less diffusion velocity of the ith species, and DTi , methodologies outlined by Bird et al. [22],
Dij, and Dij are the dimensionless thermal, mass, Hirschfelder et al. [11], and Takahashi [23].
and binary mass diffusion coefficients, respec- The governing system is discretized on a stag-
tively. These quantities are defined in a manner gered grid in generalized curvilinear coordinates.
consistent with Hirschfelder et al. [11]. Terms Xi This formulation provides non-dissipative spec-
and Yi represent respective mole and mass frac- trally clean damping characteristics and discrete
tions. The term conservation of mass, momentum and total-en-
"   # ergy. Integration is performed using a unique
XN
Xi oli
di ¼ rX j dual-time multistage scheme with a generalized
j¼1
Ru T oX j p;T ;X k ;...;ðk6¼i;jÞ all-Mach-number preconditioning methodology
j6¼i that optimally treats convective, diffusive, geomet-
þ ZðX i  Y i Þrðln pÞ ð7Þ ric, and source term anomalies in a unified man-
ner. The implicit formulation is A-stable, which
is the dimensionless form of the mechanical driv- allows one to set the physical timestep based so-
ing force, where li represents the chemical poten- lely on accuracy considerations. The scheme
tial of the ith species and p is pressure. Radiation accommodates any arbitrary equation of state,
fluxes and body forces are currently neglected. handles thermodynamic non-idealities and trans-
Eqs. (1)–(4) were solved with full finite-rate port anomalies over a wide range of conditions,
hydrogen–oxygen kinetics using the mechanism and provides full thermophysical coupling (com-
developed by Westbrook and Dryer [12] and pressible and incompressible) between processes.
Yetter et al. [13]. The property evaluation scheme The algorithm has been optimized to provide
2932 J.C. Oefelein / Proceedings of the Combustion Institute 30 (2005) 2929–2937

excellent parallel scalability attributes using a dis- Table 1


tributed multiblock domain decomposition with Injector flow, fluid, and critical properties
generalized connectivity. O2 H2
Pressure (MPa) 10.1
3. Results and discussion Temperature (K) 100 150
Density (kg/m3) 1119 15.40
Viscosity (m2/s) 1.431 · 107 3.590 · 107
The theoretical–numerical framework has been
Velocity (m/s) 30.0 125
used to perform a series of hierarchical high-fidel- Flowrate (kg/s) 421.9 · 103 27.09 · 103
ity simulations of LOX–H2 flames at supercritical Mach Number 0.0394 0.120
pressure. Both LES and DNS techniques have Reynolds Number 8.386 · 105 5.571 · 105
been applied to gain a further understanding of
Critical properties of O2: pc = 5.04 MPa (49.7 atm) and
the flow dynamics, flame structure, and associated Tc = 155 K.
property variations inherent to shear-coaxial Critical properties of H2: pc = 1.30 MPa (12.8 atm) and
injection processes. In this paper, primary empha- Tc = 33.2 K.
sis has been placed on a series of fundamental
DNS calculations performed on a sector of the velocities are 30 and 125 m/s, respectively. The
German Aerospace Center (DLR) windowed corresponding mass flowrates are 421.9 · 103
combustor operated in the high-pressure test facil- and 27.09 · 103 kg/s, and the corresponding
ity (P8) in Lampoldshausen. Mach numbers are 0.0394 and 0.120.
A schematic of the computational domain is The flow Reynolds numbers in the oxygen and
given in Fig. 2. The geometry has been non- hydrogen ducts are 8.386 · 105 and 5.571 · 105.
dimensionalized using a reference length-scale of These quantities are an order of magnitude larger
d = 4 mm, which is the physical diameter of the than that exhibited in typical laboratory-scale de-
LOX jet. The combustion chamber is cylindrical, vices at atmospheric pressures. To account for the
with a uni-element LOX–H2 shear-coaxial injector inherently smaller and wider range of scales effi-
mounted on the centerline of the face plate and a ciently (and feasibly), a quasi-2d DNS was per-
converging nozzle at the exit. The chamber and formed on a 30 cylindrical sector with periodic
nozzle assembly is 110 dimensionless units long, conditions specified in the azimuthal direction. A
the chamber diameter is 12.5 units, and the nozzle grid with 12-million hexahedral cells was con-
exit diameter is 4 units. The dimensionless inner structed in generalized coordinates. The bulk of
and outer diameters of the annular H2 jet are the cells were clustered in the near-field region just
1.150 and 1.625, respectively. The LOX-post is downstream of the injector such that the esti-
0.075 units thick and is not tapered or recessed. mated local Kolmogorov time and length-scales
The injector flow conditions and fluid proper- were resolved with an average of 20 timesteps
ties are listed in Table 1. Calculations were per- and 16 cells per respective interval. The system
formed using a reference chamber pressure of was integrated in time using a dimensionless time-
10.1 MPa (100 atm). The oxygen and hydrogen step of 0.001 based on the characteristic time of
streams are injected into the system at 100 and the incoming LOX jet, dref/Uref = 133 ls.
150 K, respectively. For this set of conditions, All wall surfaces were assumed to be hydrauli-
the hydrogen stream enters in a gaseous supercrit- cally smooth. Fully developed turbulent velocity
ical state. The LOX stream, however, enters in a profiles were generated inside the injector ducts
transcritical state that is subcritical with respect by recycling the unsteady velocity field from radial
to temperature but supercritical with respect to planes at an axial distance of 2 dimensionless units
pressure. The corresponding densities are 1119 upstream of the face-plate. The stripped fields
and 15.4 kg/m3, respectively. The mean injection were corrected to maintain the mass flow rates

Fig. 2. Schematic of the computational domain.


J.C. Oefelein / Proceedings of the Combustion Institute 30 (2005) 2929–2937 2933

specified in Table 1 and then imposed at the injec- exit. Results have been non-dimensionalized using
tor inlet 8 units upstream using a non-reflecting a reference state corresponding to the LOX jet in-
pressure condition. The walls inside the injector flow condition. Figure 4 shows the instantaneous
ducts were assumed to be adiabatic. The walls in character of the dimensionless vorticity, density,
the combustion chamber were assumed to be and temperature fields just downstream of the
300 K. A zero-gradient pressure condition was injector tip. The plot of vorticity includes OH mass
applied at the nozzle exit. The current investiga- fraction contours to mark the flame location. Com-
tion is focused exclusively on the near-field region bustion occurs at near stoichiometric conditions
just downstream of the injector. This region is de- and produces a wake that effectively separates the
picted in Fig. 2 by the 15 · 4.875 rectangle adja- hydrogen and oxygen streams as the flow evolves
cent to the injector face-plate. Figure 3 shows downstream. Because of the liquid-like characteris-
contours of the normalized turbulent kinetic en- tics of the LOX, an extremely large density gradient
ergy field in this region to highlight the complex exists within the shearlayer. The close proximity of
flow structure. This result was taken from a com- the dense LOX core on one side of the flame and ex-
panion LES calculation using a full three-dimen- tremely low density fluid on the other side promotes
sional grid with 8.8-million cells. the formation of small-scale vortical structures
Analysis of the subsequent DNS results is fo- within the shearlayer that evolve dynamically in a
cused on the region of high shear near the injector manner analogous to that produced by a backward
facing step. These structures emanate from the
upper corner of the LOX-post, are initially much
smaller than the LOX-post thickness, and coalesce
within the resultant downstream wake.
The degree to which the properties vary across
the flame and shearlayer in the near-field region
are shown in Figs. 5–9. Symbols mark the location
of respective grid nodes. Figure 5 shows the radial
Fig. 3. Normalized turbulent kinetic energy field.

Fig. 5. Radial variation in the Kolmogorov time and


length scales at an axial location of 0.03 units.

Fig. 4. Instantaneous (A) vorticity, (B) density, and (C) Fig. 6. Radial variation of temperature and density at
temperature fields. an axial location of 0.03 units.
2934 J.C. Oefelein / Proceedings of the Combustion Institute 30 (2005) 2929–2937

Fig. 7. Radial variation of H2, O2, OH, and H2O mass


fractions at an axial location of 0.03 units.

Fig. 8. Radial variation of Z and derivative relations Zp


and ZT at an axial location of 0.03 units.

variation in the dimensionless Kolmogorov time


and length scales at an axial location of 0.03 units.
The Kolmogorov time-scales approach values
as small as 102 in the flame and 103 on the hydro-
gen side of the shearlayer. The Kolmogorov length-
scales approach values as small as 104 and 105,
respectively. This figure also shows how the grid
spacing distribution compares to the local Kol-
mogorov length-scales across the shearlayer to ver-
ify the degree of resolution achieved in the
calculation. The grid spacing in the axial direction Fig. 9. Radial variation of kinematic viscosity, Prandtl,
provides a similar degree of resolution. Schmidt, and Lewis numbers at an axial location of 0.03
Figures 6 and 7 show the corresponding radial units.
variations in temperature, density, and the H2, O2,
OH, and H2O mass fractions. The change in den- files of the compressibility factor Z and the
sity across the flame is of the order of 1000 to 1 derivative relations
   
and occurs over an interval that is only a fraction p oZ
of the LOX-post thickness. An equally distinct Zp ¼ 1  ð8Þ
Z op T
change in temperature due to combustion and sca-
lar-mixing is also evident. Though continuous, and
"  #
gradients in the flame zone approach the behavior T oZ
of a contact discontinuity. Depletion of all of the ZT ¼ 1 þ ð9Þ
Z oT p
oxygen in the flame, the flame structure, and the
fuel-rich characteristics on the hydrogen side of are shown. The compressibility factor indicates
the flame are also apparent. the degree to which the local mixture state ap-
Figure 8 provides a further characterization of proaches that of an ideal gas. Eqs. (8) and (9)
the thermodynamic properties across the shear- are directly related to the isothermal compressibil-
layer at the same axial location. Here, radial pro- ity and the coefficient of thermal expansion and
J.C. Oefelein / Proceedings of the Combustion Institute 30 (2005) 2929–2937 2935

indicate the degree to which these fundamental stagnation region that provides the primary
forms of compressibility deviate from that of an flame-holding mechanism. The fluid dynamic as-
ideal gas. The isothermal compressibility charac- pects of this mechanism are shown in Fig. 10.
terizes the change in volume that results from a Flame development in the stagnation region is
change in pressure while the temperature remains dominated by diffusion and enhanced by convec-
constant. The coefficient of thermal expansion tive energy transport from the backflow of fuel-
characterizes the change in volume that results rich combustion products. The backflow of hot
from a change in temperature while the pressure products sustains ignition and promotes vaporiza-
remains constant. Figure 8 shows that real-gas tion and mixing of the dense LOX core.
and liquid effects exist locally in colder regions of
the flow on either side of the shearlayer, and ideal
gas conditions exist within the hot fuel-rich region 4. Conclusions
and flame zone. This latter observation can be di-
rectly attributed to the inherent diminished effect The thermophysical characteristics of a liquid-
of pressure with increasing temperature. oxygen–hydrogen flame at supercritical pressure
Figure 9 shows the radial variation in the have been investigated using the direct numerical
dimensionless kinematic viscosity and the Fickian simulation technique. Calculations were performed
(mixture-averaged) values of the Prandtl, Schmidt, on a sector of a computational domain designed to
and Lewis numbers, respectively. Relative to the match the German Aerospace Center (DLR) win-
value in the LOX core, the kinematic viscosity in- dowed combustor operated in the high-pressure
creases by a factor of approximately 100 in the test facility (P8) in Lampoldshausen. The combus-
flame zone and varies by a factor of 10–100 in the tion chamber is cylindrical, with a uni-element
shearlayer. The Prandtl, Schmidt, and Lewis num- shear-coaxial injector mounted on the centerline
bers all decrease from values greater than 1 on the of the face-plate and a converging nozzle at the exit.
LOX side of the flame to values less than 1 on the Oxygen was injected in a cryogenic state at a sub-
hydrogen side. Relative magnitudes on the LOX critical temperature and supercritical pressure.
side suggest that the diffusion of momentum domi- The hydrogen was injected in a pure supercritical
nates over energy and mass, and that the diffusion state. Emphasis was placed on the near-field region
of energy dominates over mass. Mass diffusion just downstream of the injector element. In this re-
rates into the flame on the LOX side are thus rate- gion, a diffusion dominated mode of combustion
limiting. Values associated with the fuel-rich prod- exists in the presence of exceedingly large thermo-
ucts, on the other hand, suggest that the opposite physical property gradients. The flame anchors it-
trend will be present. On this side of the flame, the self within the interfacial region of high shear that
relative magnitudes of the Prandtl, Schmidt, and exists between the liquid-oxygen core and the annu-
Lewis numbers suggest that the diffusion of energy lar hydrogen jet. Resultant variations in the ther-
will dominate over momentum, and that mass diffu- mophysical properties across the flame intensify
sion dominates over both momentum and energy. respective gradients such that they approach the
The results presented above suggest that the behavior of a contact discontinuity.
smoothness of the LOX jet can be attributed to Results provide detailed quantitative character-
high axial inertia coupled with rapid radial heat- izations of the flame structure, associated property
ing and vaporization within the shearlayer. The variations, and the dominating effects of the dense
initial vortical expansion of the low-density liquid-oxygen core on the near-field shearlayer
hydrogen induces strong recirculating backflow dynamics. Diminished mass diffusion rates inside
in the vicinity of the LOX-post. The net effect of the dense core create a fuel-rich condition within
the local thermophysical interactions is to produce the near-field shearlayer region. Combustion oc-
a strong counter-recirculating zone of hot fuel- curs at near stoichiometric conditions and produces
rich products in the vicinity of the LOX-post a wake that effectively separates the hydrogen and
tip. Inside this recirculation zone is an unsteady oxygen streams as the flow evolves downstream.
The extremely low values of density in the hot reac-
tion zone cause a considerable reduction of aerody-
namic interactions between the liquid-oxygen jet
and the hydrogen stream. Significant real-gas and
liquid behavior is observed to coexist locally in
colder regions of the flow, with ideal gas and trans-
port characteristics existing within the reaction and
product zones. The latter can be directly attributed
to the inherent diminished effect of pressure with
increasing temperature.
The coupled effect of local thermophysical
interactions is to produce a strong counter-
Fig. 10. Flame-holding mechanism. recirculating zone of hot fuel-rich products in
2936 J.C. Oefelein / Proceedings of the Combustion Institute 30 (2005) 2929–2937

the vicinity of the injector tip. This recirculation [8] O. Gurliat, V. Schmidt, O.J. Haidn, M. Oschwald,
zone provides the primary flame-holding mecha- Aerospace Sci. Technol. 7 (2003) 517–531.
nism. The flame-holding mechanism is dominated [9] J.C. Oefelein, General numerical framework
by an unsteady stagnation region. Flame develop- for reacting multiphase flow with complex
thermochemistry, thermodynamics and trans-
ment within the stagnation region is dominated by port, copyright 1992–2004 All Rights Reserved
diffusion and enhanced by convective energy (2004).
transport from the backflow of fuel-rich combus- [10] J.C. Oefelein, General package for evaluation of
tion products. The backflow of hot products both multicomponent real-gas and liquid mixture states
sustains the ignition process and promotes vapor- at all pressures, copyright 1992–2004 All Rights
ization of the dense liquid-oxygen core. Reserved (2004).
[11] J.O. Hirschfelder, C.F. Curtiss, R.B. Bird, Molec-
ular Theory of Gases and Liquids, second ed., Wiley,
New York, 1964.
Acknowledgment [12] C.K. Westbrook, F.L. Dryer, Prog. Energy Com-
bust. Sci. 10 (1) (1984) 1–57.
This work was funded by the NASA Marshall [13] R.A. Yetter, F.L. Dryer, H. Rabitz, Combust. Sci.
Space Flight Center under Contract No. H- Technol. 79 (1991) 97–128.
33477D. The support provided is gratefully [14] T.W. Leland, P.S. Chappelear, Ind. Eng. Chem.
acknowledged. Fundam. 60 (7) (1968) 15–43.
[15] J.S. Rowlinson, I.D. Watson, Chem. Eng. Sci. 24 (8)
(1969) 1565–1574.
[16] R.C. Reid, J.M. Prausnitz, B.E. Polling, The Prop-
References erties of Liquids and Gases, fourth ed., McGraw-
Hill, New York, 1987.
[1] W. Mayer, H. Tamura, J. Propul. Power 12 (6) [17] J.C. Oefelein, Simulation and Analysis of Turbulent
(1996) 1137–1147. Multiphase Combustion Processes at High Pressures,
[2] W. Mayer, A. Schik, B. Vieille, C. Chaveau, I. Ph.D. thesis, The Pennsylvania State University,
Gökalp, D. Talley, R. Woodward, J. Propul. Power University Park, Pennsylvania, 1997.
14 (5) (1998) 835–842. [18] J.C. Oefelein, V. Yang, J. Propul. Power 14 (5)
[3] W. Mayer, A. Schik, M. Schäffler, H. Tamura, J. (1998) 843–857.
Propul. Power 16 (5) (2000) 823–828. [19] G.J. VanWylen, R.E. Sonntag, Fundamentals of
[4] G. Herding, R. Snyder, C. Rolon, S. Candel, J. Classical Thermodynamics, third ed., Wiley, New
Propul. Power 13 (2) (1998) 146–151. York, 1986.
[5] S. Candel, G. Herding, R. Snyder, P. Scouflaire, C. [20] J.F. Ely, H.J.M. Hanley, Ind. Eng. Chem. Fundam.
Rolon, L. Vingert, J. Propul. Power 14 (5) (1998) 20 (4) (1981) 323–332.
826–834. [21] J.F. Ely, H.J.M. Hanley, Ind. Eng. Chem. Fundam.
[6] D. Kendrick, G. Herding, P. Scouflaire, J.C. Rolon, 22 (1) (1981) 90–97.
S. Candel, Combust. Flame 118 (1999) 327–339. [22] R.B. Bird, W.E. Stewart, E.N. Lightfoot, Transport
[7] M. Juniper, A. Tripathi, P. Scouflaire, J.C. Rolon, Phenomena, Wiley, New York, 1960.
S. Candel, Proc. Combust. Inst. 28 (1) (2000) [23] S. Takahashi, J. Chem. Eng. Jpn. 7 (6) (1974)
1103–1109. 417–420.

Comment

Sebastien Candel, EM2C Lab, CNRS, Ecole Centale Reply. Differences between the current simulation
Paris, France. Your calculations are quite close to some (see Fig. 4) and data gathered from recent cryogenic
of the data gathered in cryogenic flame experiments car- flame experiments can be directly attributed to the man-
ried out in recent years [1] but there is also one differ- ner in which the wall boundary conditions are specified.
ence. Your flame is attached at the point where the This is an extremely important and valid observation.
oxygen leaves the injector. In our experiments, the flame Results from the simulation are based on the assumption
has an edge, and it is anchored at a small but finite dis- that the injector and faceplate walls are adiabatic. This
tance from the injection lip. assumption produces a much higher wall temperature
You also mentioned that some extinction takes place in the vicinity of the flame than would be observed in
at a distance from the injection plane. We found that the actual experiment and does not accommodate
oxygen/hydrogen flames require excessively large strain quenching in the vicinity of the wall. In reality, the flux
rates to be extinguished. Is there a possibility that this distribution across the walls is quite complex, particu-
extinction could be related to an insufficient resolution larly in the vicinity of the LOX-post tip. Enhanced
of the flame reaction layer? heat-transfer within the walls will tend to both smooth
and reduce the wall temperature distribution to values
Reference much closer to the propellant injection temperatures of
100 and 150 K (see Table 1). Detailed treatment of wall
[1] M. Juniper, A. Tripathi, P. Scouflaire, J.-C. Rolon, effects is clearly an important factor that must eventually
S. Candel, Proc. Comb. Inst. 28 (2000) 1103–1109. be incorporated into these types of simulations.
J.C. Oefelein / Proceedings of the Combustion Institute 30 (2005) 2929–2937 2937

Subsequent analysis of the coupled flow-flame interac- the reaction-rate associated with the production of OH
tions has suggested that the vortical structures emanating radicals was observed to drop significantly by a factor of
from the hydrogen side of the LOX-post can produce a rel- 1–2 orders of magnitude. Within this region, the OH pro-
atively strong dilatation field. This field coalesces progres- duction rate profiles are resolved with approximately 12
sively within the LOX-post wake and subsequently cells in each spatial direction and 8–16 time-steps locally
interacts with the flame a short distance downstream. In as the drop in production rate occurs. This is a distinctly
regions where the dilatation field compresses the flame interesting observation that requires further investigation.

Vous aimerez peut-être aussi