You are on page 1of 8

Proceedings

of the
Combustion
Institute
Proceedings of the Combustion Institute 31 (2007) 2497–2504
www.elsevier.com/locate/proci

LES of supersonic combustion in a scramjet


engine model
a,b,* a
M. Berglund , C. Fureby
a
Weapons and Protection Division, Swedish Defence Research Agency (FOI), SE–164 90 Stockholm, Sweden
b
Division of Fluid Mechanics, Department of Energy Sciences, Lund Institute of Technology, Box 118,
SE–221 00 Lund, Sweden

Abstract

In this study, Large Eddy Simulation (LES) has been used to examine supersonic flow and combustion
in a model scramjet combustor. The LES model is based on an unstructured finite volume discretization,
using total variational diminishing flux reconstruction, of the filtered continuity, momentum, enthalpy, and
passive/reactive scalar equations, used to describe the combustion process. The configuration used is sim-
ilar to the laboratory scramjet at the Institute for Chemical Propulsion of the German Aerospace Center
(DLR) and consists of a one-sided divergent channel with a wedge-shaped flameholder at the base of which
hydrogen is injected. Here, we investigate supersonic flow with hydrogen injection and supersonic flow with
hydrogen injection and combustion. For the purpose of validation, the LES results are compared with
experimental data for velocity and temperature at different cross-sections. In addition, qualitative compar-
isons are also made between predicted and measured shadowgraph images. The LES computations are
capable of predicting both the non-reacting and reacting flowfields reasonably well—in particular we notice
that the LES model identifies and differentiates between peculiarities of the flowfields found in the
experiments.
 2006 The Combustion Institute. Published by Elsevier Inc. All rights reserved.

Keywords: Large Eddy Simulation; Flamelet model; Supersonic combustion; Scramjet

1. Introduction an efficient propulsion system. Turbojets and rock-


et engines, used for high-speed aircraft and space
The current road-map of aerospace technology launcher and missile systems, respectively, are
development contains several areas of application unattractive from technical, economical as well as
for hypersonic flight vehicles such as long-range safety points-of-view. The new pulse detonation
passenger transport, reusable launch vehicles for technology may hold promise for the future but
space applications, and long-range missiles. The needs to mature. The preferred propulsion system
successful development of such vehicles would for flight in the supersonic (3 < Ma < 5) regime is
depend, to a large extent, on the development of a ramjet and beyond that, in the hypersonic
(5 < Ma < 15) regime, is a ramjet with supersonic
combustion, i.e. a scramjet. Due to the supersonic
*
Corresponding author. Fax: +46 8 5550 4144. flow speed (and hence short residence times) in
E-mail address: magnus.berglund@foi.se (M. the combustor of a scramjet, problems arise in the
Berglund). mixing of fuel and oxidizer to form a combustible

1540-7489/$ - see front matter  2006 The Combustion Institute. Published by Elsevier Inc. All rights reserved.
doi:10.1016/j.proci.2006.07.074
2498 M. Berglund, C. Fureby / Proceedings of the Combustion Institute 31 (2007) 2497–2504

mixture, flame anchoring and stability, thermal the complex flow physics and chemistry in this
choking due to uncontrolled heat release, and com- simplified scramjet combustor. Different resolu-
pletion of combustion within a finite length com- tions have been used and the similarity of the pre-
bustor. A further complication is that the dictions and the experiments suggests that the
scramjet combustion experiments are very compli- physical processes are reasonably well described.
cated, and only a few limited runtime facilities are Previous computational studies of this laboratory
available. The most cost-effective way of investigat- scramjet engine includes a RANS-based study [10]
ing scramjet combustion therefore lies in the use of and two LES-based studies [11,12].
computational fluid dynamics (CFD), provided
that the models have the required fidelity.
In high-speed turbulent combustion, density 2. LES modeling of supersonic combustion
variations arise not only from exothermicity but
also from viscous heating, compression, and The governing equations for turbulent reacting
expansions associated with the high-speed. The flows are the reactive Navier–Stokes Equations
quantity that distinguishes these flows from low- (NSE). Due to the variety of scales, in particular
speed flows is the dilatation, $ Æ v. In low-speed the disparity of flow and chemistry scales, and
turbulent combustion the influence of pressure the large number of species involved in combus-
and velocity on $ Æ v can virtually be neglected, tion of practical fuels, these equations must be
and under these circumstances $ Æ v is due only simplified. In LES one usually choose between:
to the exothermicity, whereas in high-speed com- (i) reduced or global reaction mechanisms with
bustion all of the above features play a role. As Arrhenius rate expressions in the framework of
the speed increases the kinetic energy dominates Implicit LES [13], (ii) flamelet models, in which
the enthalpy so that the exothermicity adds rela- reaction is assumed to occur in thin layers, (iii) lin-
tively smaller amounts of energy, and accordingly, ear eddy models, using a grid-within-the-grid
the influence of combustion becomes less impor- method to solve 1D species equations with full
tant. The description of chemical kinetics is com- resolution, (iv) thickened flame models (TFM),
plicated, involving either a large number of species and (v) probability density function (PDF) meth-
or extensive simplifications resulting in reduced ods. See Poinsot and Veynante [1] for a review.
reaction mechanisms. Moreover, chemical kinetics Bray [14] argues that the most probable range
and species diffusion influence each other and for scramjet combustion is within the flamelet
occur mainly on micro-scales together with mix- regime, which is further supported by an analysis
ing, counter-gradient diffusion, flame-generated by Waidman et al. [9], based on experimental data
turbulence, and various types of instabilities. from the laboratory scramjet engine used here.
A promising method for predicting scramjet Based on the LES results the ratio between the
combustion is Large Eddy Simulation (LES) [1–3] integral length scale and the laminar flame thick-
in which the largest turbulence scales are explicitly ness is estimated to be 200 and the ratio between
computed and only the small, low-energy modes the rms velocity and the laminar flame speed to be
are modeled (where the assumption of local isotro- 5, so that Ka  1, Da  40, and Re  1. This
py is expected to hold better than for the largest flow case is thus in the corrugated flamelet regime.
scales). The direct computation of the large energy In essence, the flamelet assumption allows decou-
containing eddies (which are geometry and flow pling of the chemistry from the fluid mechanics
dependent) gives LES more generality than Rey- and the scale separation enables pre-computing
nolds Averaged Navier–Stokes (RANS) [1,4] in the chemistry as parameterized by one or more
which all turbulent scales are modeled. At present, passive and/or reactive scalars. Two flamelet mod-
the use of LES to investigate non-reacting and sub- els of different fidelity will be used here in conjunc-
sonic reacting flows is becoming established tion with the two-step global reaction mechanism
whereas the application of LES to supersonic react- of Rogers and Chinitz [15].
ing flows has so far been limited. As for low-speed
flows, mixing and chemical kinetics occur mainly 2.1. One-equation flamelet model
on the subgrid level, and will require modeling [5].
In the present study, the experimental work of In the one-equation flamelet model the filtered
Waidmann et al. [6–9] is used as reference and to NSE [16] are solved together with a passive scalar
provide further validation of the supersonic LES (here, we use the mixture fraction) used to param-
combustion model developed and used here. The eterize the chemistry,
laboratory test rig consists of a rectilinear com-
bustion chamber with a slanted upper wall and a ot ðqÞ þ r  ðq~vÞ ¼ 0
wedge-shaped flameholder at the base of which
gaseous hydrogen, H2, is injected. Here, we use ot ðq~vÞ þ r  ðq~v~vÞ ¼ rp þ r  ðS  BÞ
two flamelet models of different fidelity to exam- ð1Þ
~ þ r  ðq~vEÞ
ot ðqEÞ ~ v þ h  bE Þ
~ ¼ r  ððp1 þ SÞ~
ine the supersonic mixing and combustion taking
place downstream of the wedge, and to explore ot ðq~zÞ þ r  ðq~v~zÞ ¼ r  ðDz r~zÞ  r  bz
M. Berglund, C. Fureby / Proceedings of the Combustion Institute 31 (2007) 2497–2504 2499

where q is the density, v the velocity, p the pres- displacement term describing the propagation of
sure, S the viscous stress tensor, B the subgrid f at the laminar flame speed Su = Su(z), i.e.
stress tensor, E ¼ h  p=q þ 12 v2 the total ener- r  ðqDrfÞ  x_ f ¼ qS u jrfj. The flame-front
gy (where h is the mixture enthalpy), h the heat displacement term can be expanded as
flux vector, bE the subgrid flux vector of E, z the qS u jrfj ¼ hqS u iR ¼ hqS u iNjrfj, where R ¼ jrfj
mixture fraction, and bz the subgrid flux vector
is the flame surface density, N ¼ R=jrfj the
of z. The subgrid stress tensor B and subgrid flux
vectors bE and bz are here closed using the mixed flame-wrinkling, and hqS u i ¼ qS u jrfj=R is the
model [17] in which surface filtered consumption rate, so that the fil-
tered transport equation for ~f finally becomes
~ D;
B ¼ qð~v~v  ~v~vÞ þ 23qkI  2lk D
ot ðq~fÞ þ r  ðq~v~fÞ ¼ r  bf  hqS u iNjrfj: ð4Þ
bE ¼ qð~vE ~  lk rE;
~  ~vEÞ ~ ð2Þ
bz ¼ qð~v~z  ~v~zÞ  lk r~z; In order to close Eq. (4) sub-models for bf, N, and
ÆqSuæ must be provided. Following Fureby [21] the
where the subgrid kinetic energy and viscosity are fractal flame-wrinkling (FFW) model is used for N
modeled by k ¼ cI D2 kDk ~ 2 and lk ¼ qcD D2 kDk,~ together with the mixed model for bz and bf. In
~
respectively, where D is the rate-of-strain tensor the FFW model N is interpreted as the ratio be-
and the subscript D denotes the deviatoric part. tween the wrinkled and undisturbed flame surface,
The species mass fractions, Yi, are assumed N = (eo/ei)D2, where ei the inner cut-off, and D
to be functions of z and the dissipation rate, v, the fractal dimension. In [21] ei is defined as the
such that Yi = Yi(z,v). Then, the filtered mass subgrid wrinkling length scale that is here evaluat-
Rfractions, Y~ i , can be defined by Y~ i ¼ ed from the curvature tensor, so that
1 R1
0 0
Y i ðz; vÞpðz; vÞ dz dv, where p(z,v) is the sub- ei ¼ S 0u D=v0 C, where C ¼ CðD=d0u ; v0 =S 0u Þ is the effi-
grid probability density function (PDF) [18]. ciency function [22], v 0 the subgrid velocity, d0u the
Assuming that z and v are uncorrelated and that laminar flame thickness, and S 0u the unstrained
the effects of the subgrid fluctuations in v can be laminar flame speed so that N ¼ ðCv0 =S 0u ÞD2 . Fol-
neglected pðz; vÞ ¼ pðzÞdðv  ~vÞ, where d is the lowing [23], the fractal dimension is modeled by
Dirac function. We further assume a Beta distri- D ¼ 2:00=ðv0 =S 0u þ 1Þ þ 2:35=ðS 0u =v0 þ 1Þ and we
bution for z, parameterized by ~z and its variance further assume that the inner structure of the reac-
ð~z00 Þ2  cz D2 kr~zk2 . In addition, ~v is modeled by tion zone is unaffected by the turbulence so that
v ¼ 2cv DkDkð~
~ ~ z00 Þ2 . The model coefficients are ob- hqS u i  qu S u ð~zÞ. The thermodynamic model and
tained from an inertial subrange behavior the diffusive transport properties are included in
[17,19]. The flamelet library includes H2, O2, a manner similar to the one-equation flamelet
OH, H2O, and N2 [15], and is tabulated for differ- model.
ent values of z and v. The thermodynamic model
consists of the equation-of-state for an ideal gas
mixture and the equation-of-state for the enthalpy 3. Numerical methods
of the mixture. Diffusive transport properties
(using a Fickian scheme for ~z), mixture viscosity A finite volume (FV) method for arbitrary cell-
(using Sutherland’s law), and thermal conduction shapes together with an explicit, fully compress-
(modeled similarly) are included, whilst thermal ible solver is used to solve the governing equations
diffusion is neglected. for the reactive LES models. The code employs a
colocated cell-centered variable arrangement and
2.2. Two-equation flamelet model the semi-discretized LES equations are obtained
from Gauss theorem. To complete the FV-discret-
In the two-equation flamelet model an attempt ization the fluxes need to be reconstructed from
is made to resolve the limitations of the one-equa- variables at adjacent cells. High-order reconstruc-
tion flamelet model, which are caused by the lack tion of all convective fluxes, using a second order
of information about the chemical reactions and (flux limiter-based) TVD scheme and central dif-
the chemical variations perpendicular to r~z. The ferencing of the inner derivatives in the viscous
remedy is to use both a passive scalar, ~z, and a and subgrid fluxes gives nominally second order
reactive scalar, ~f, so that the model consists of accuracy in space. The time-integration is fully
Eq. (1) supplemented with explicit using the second order accurate TVD
Runge–Kutta scheme of Gottlieb and Shu [24].
ot ðq~fÞ þ r  ðq~v~fÞ ¼ r  ðDf r~fÞ  r  bf  x_ f : The governing equations are thus solved with full
ð3Þ coupling between the equations. An issue for LES
is how the subgrid model interacts with the lead-
Following Duwig and Fureby [20] the reaction ing order truncation error. In [25] this aspect has
rate, x_ f , and diffusive term on the right hand side been studied, in the context of implicit LES, using
of Eq. (3) are regrouped as a filtered flame-front modified equation analysis. These results are
2500 M. Berglund, C. Fureby / Proceedings of the Combustion Institute 31 (2007) 2497–2504

extendable to reacting LES and for the present The computational configuration include three
model we find that the subgrid model, e.g. holes together with imposed periodicity in the
~ D , and the leading
B ¼ qð~v~v  ~v ~vÞ þ 23 qkI  2lk D spanwise direction. Two grids with approximately
order truncation error, e.g. T ¼ qðC  ðr~vÞT þ 3.2 and 6.4 million cells are used for the calcula-
ðr~vÞ  CT þ v2 ðr~v  dÞðr~v  dÞÞ, where C and tions. The grids are clustered towards the upper,
v are functions of the flux limiter C, may overlap lower, and wedge walls, as well as in the wake
in regions with large gradients. The scale similar- region and the shear layer region to resolve the
ity term in the subgrid model is, however, the mixing. There are no significant differences
dominant term and since C „ 0 only in highly between the results for first and second order sta-
localized regions we do not expect the leading tistical moments on the two meshes.
order truncation error to have significant impact Dirichlet boundary conditions are used for all
on the solutions. variables at the inlet and at the H2-jets at the base
of the strut. At the outlet, all variable values are
extrapolated from the interior. At the upper, low-
4. Experimental setup and computational details er, and strut walls, zero Dirichlet conditions are
applied to the velocity whereas zero Neumann
A schematic of the scramjet experimental facil- conditions are applied to all other variables.
ity, [6–9], is presented in Fig. 1. Preheated air is To handle the near-wall resolution problem
expanded through a laval nozzle and enters the subgrid wall-models [26], in which lk is modified
combustor section at Ma = 2.0. The combustor for the flow to satisfy a logarithmic velocity distri-
has a width of 40 mm and a height of 50 mm at bution, are used.
the entrance and a divergence angle of the upper All computations are initialized with the state
channel wall of three degrees to compensate for of the incoming air and are continued until the
the expansion of the boundary layer. A wedge- second order statistical moments have converged;
shaped strut is placed in the combustion chamber after about five flow through-times.
downstream of the nozzle. Just downstream of the
nozzle the height of the 32 mm long strut is 6 mm.
Along the first 10 cm downstream of the nozzle, 5. Results and discussion
the side walls and the upper wall are made from
quartz glass to allow optical access and to mini- In [12] the flow without and with H2 injection
mize the reflection of scattered light on the wall in the scramjet engine model was examined using
opposite the observation window. Hydrogen LES and here we continue this study by further
(H2) is injected at Ma = 1.0 through a row of 15 examining the case with H2 injection and with
holes, 1.0 mm in diameter and 2.4 mm apart, in H2 injection and combustion. Inert H2 injection
the strut base. Typical mass flows in the experi- adds significant complexity to the flow in the
ments were varied between 1.0 and 1.5 kg/s for scramjet combustor since H2 has a considerably
the air and between 1.5 and 4.0 g/s for H2, which lower molar mass than air, which makes mixing
correspond to equivalence ratios between 0.034 an important process in establishing the condi-
and 0.136, respectively. The hydrogen is injected tions for scramjet combustion.
at ambient temperature and pressure, i.e. at Figure 2 presents a perspective view of the flow
T = 250 K and p = 105 Pa, whereas the air was together with a qualitative comparison of experi-
injected at T = 340 K and p = 105 Pa. Combus- mental and numerical shadowgraph images from
tion was initiated by pre-burning of a small the cases with inert H2 injection (without combus-
amount of O2 in a H2 tube by a spark. The exper- tion). With inert H2 injection, oblique shocks are
iments include LDV and PIV measurements of formed at the tip of the wedge that are later reflect-
velocity, CARS measurements of temperature, ed by the upper and lower walls before interacting
OH-LIF for mapping regions of combustion, further with the unsteady, partly H2 filled, wake.
and schlieren and shadowgraph photography in Together with the slightly bent expansion fan com-
order to characterize the flow and combustion ing off the base of the wedge this causes a charac-
dynamics. teristic shock-wave pattern in the downstream
region. At the upper and lower walls, the boundary
layer is affected, at least locally, by the reflected
100 (mm) oblique shocks. These local modifications involve
thickening of the boundary layer, increased rms-
50
air, T=340 K hydrogen, T=250 K pressure fluctuations, and elevated wall tempera-
Ma=2.0 Ma=1.0
25
tures. With the exception of the reflected shocks
109 interacting with and further reflecting off the partly
340 H2 filled wake, these structures are similar to the
ones found without H2 injection, cf. [8,9,12]. The
Fig. 1. Schematic of the supersonic combustion cham- boundary layer on the wedge surface separates at
ber and the coordinate system used. the base and a shear layer is formed. This shear
M. Berglund, C. Fureby / Proceedings of the Combustion Institute 31 (2007) 2497–2504 2501

shocks (reflecting off the walls) pass through the


accelerating wake. The resemblance between the
predicted and the experimental shadowgraph
images is good.
Figure 3 shows a perspective view of the react-
ing flow together with a qualitative comparison of
experimental and numerical shadowgraph images
from the case with H2 injection (at 1.50 g/s) fol-
lowed by combustion. With H2 injection and com-
bustion the expansion fans at the upper and lower
corners of the wedge essentially vanish, whereas
the recompression shocks become weaker as com-
pared to the inert H2 injection case. With combus-
tion the recirculation region becomes longer and
wider, and serves as a flameholder for the H2 dif-
fusion flame. Moreover, from the LES, the peak
reverse velocity (in the recirculation region) and
the base pressure increase as compared to the inert

Fig. 2. Non-reacting case: (a) perspective view in terms


of contours of the axial momentum on the back plane
and an iso-surface of the H2 mass fraction; (b) exper-
imental shadowgraph; and (c) numerical shadowgraph
(essentially r2 q) image.

layer is naturally unstable and is therefore prone to


break-up and develops into Kelvin–Helmholtz
(KH) structures. Because of the one-sided diver-
gent channel the upper reflecting shock hits the
H2 filled wake further downstream than the lower
shock, causing an asymmetric flow field through
which the KH modes are amplified. This, in turn,
results in shedding and periodically bent structures
whereby air and H2 are mixed by advection. By
comparing the LES results with the Rayleigh scat-
tering images in [8,9] (not shown) we find very
good agreement for the mixing pattern. It is impor-
tant also to notice that compressibility leads to
reduced shear layer growth, affecting the forma-
Fig. 3. Reacting case: (a) perspective view in terms of
tion and subsequent break-up of large coherent contours of the axial momentum on the back plane and
structures. In addition, the reflected shock waves on the two cross planes, contours of the temperature, an
are deflected by the hydrogen jets. After some dis- iso-surface of the flame and an iso-surface of the H2
tance the flow in the wake of the wedge is acceler- mass fraction; (b) experimental shadowgraph; and
ated back to supersonic speed and the subsequent (c) numerical shadowgraph (essentially r2 q) image.
2502 M. Berglund, C. Fureby / Proceedings of the Combustion Institute 31 (2007) 2497–2504

H2 injection case, which is supported by the exper- a x/D=22.00 x/D=34.50


x/D=20.83
iments [8]. The shear layers, originating at the 0.05
x/D=13.00

upper and lower corners of the wedge, are much x/D=26.16

more pronounced in the combusting case because 0.04

ignition occurs within the shear layers between the

y [m]
0.03
H2 rich wake and the free-stream air. The flow can
roughly be divided into three regions: the induc- 0.02

tion zone, where the turbulence determines the 0.01


mixing and the progress of combustion; a transi-
tional zone that is dominated by large-scale coher- 0
0 500 1000 1500 2000 2500
ent flow structures, convective mixing, air vx [m/s]

entrainment and exothermicity, and a turbulent


combustion zone, with large-scale coherent struc- b x/D=22.00
x/D=20.83
0.05 x/D=13.00 x/D=34.50
tures and mixing. The large-scale structures origi- x/D=26.16
nate in the shear layers that roll up, and become 0.04

increasingly distorted with downstream distance

y [m]
0.03
due to vortex break-down, occupying a large part
of the combustor due to the dilatation, resulting 0.02

both from the flow and the exothermicity. These


0.01
structures are formed by the interaction of the
shocks with the unstable shear layers. The exo- 0
0 200 400 600 800
thermicity increases the shear layer thickness vrms [m/s]
x

which, in turn, affects the subsequent reflections


of the oblique shocks and the overall combustion c
x/D=20.83 x/D=34.50
chamber pressure. The large-scale structures cause 0.05 x/D=13.00
x/D=26.16
momentum exchange by mixing cold high-mo- 0.04
mentum air with hot low-momentum H2 or prod-
y [m]

uct wake flow enforcing the transition from 0.03

subsonic wake flow to supersonic free-stream flow 0.02


together with exothermicity within these struc-
tures. Concerning the agreement between predict- 0.01

ed and experimental shadowgraph images we find 0


0 1000 2000 3000 4000 5000 6000
only fair agreement; the computed images, here T [K]
and in [11], both underestimate the growth of
the wake resulting from the volumetric expansion Fig. 4. (a) Time-averaged axial velocity, (b) axial rms-
due to the reactions in the transitional region just velocity fluctuations, and (c) temperature at different
downstream of the wedge. The reason(s) for this cross-sections downstream of the wedge. Legend: Gray
are not fully understood at present, but may be curves and symbols (·) denote the LES and exp. data,
respectively, for the inert H2 injection case whereas black
related to (i) poor subgrid mixing models, (ii) lines (solid for the two-equation model and dashed for
the common use of the two-step reaction mecha- the one-equation model) and symbols (+) denote LES
nism [15] which may be too simple, (iii) the use and exp. data, respectively, for the combusting case.
of a calorically perfect gas model instead of a ther- Note also that profiles are shifted in the horizontal
mally perfect gas model, and (iv) inaccuracies direction.
related to the modeling of the shadowgraph imag-
es as r2 q. On the other hand, the present LES observed on each side of the hydrogen jet(s), and
results are supported by the OH-LIF and Ray- reasonable agreement with the rather sparse exper-
leigh scattering results in [9] where the effect of imental h~vx i profile is found. At the second station
the volumetric expansion is substantially less pro- (x/h = 20.83) the predicted h~vx i defect is narrower
nounced as compared to the experimental shad- than the measured one, which also shows a pecu-
owgraph image. No reasons for this somewhat liar asymmetric shape. Previous RANS calcula-
contradictory results from the experiments are tions [10] and the LES of Génin et al. [11] show
known to the authors. More detailed comparisons very similar profiles as presented here, and at pres-
between OH-LIF and Rayleigh scattering results ent the origin of the discrepancy between the pre-
from experiments and LES data will be reported dictions and the measurement data is unclear. At
elsewhere [27]. the third and fourth station (x/h = 22.00 and
In Fig. 4a we show profiles of the time-averaged 26.16) good agreement is again obtained, for both
axial velocity, h~vx i, at x/h = 13.00, 20.83, 22.00, the inert mixing and combustion cases, now result-
26.16, and 34.50 along with experimental data ing in an almost flat h~vx i profile with a constant
when available for both the inert H2 injection case velocity of about 760 m/s that remains flat
and the reacting case. At the first cross-section throughout the computational domain. This
(x/h = 13.00) strong reverse velocity may be should be contrasted against the 2D RANS
M. Berglund, C. Fureby / Proceedings of the Combustion Institute 31 (2007) 2497–2504 2503

results, predicting a strong overshoot in h~vx i at x/ differencing for the viscous fluxes. The time-inte-
h = 20.83, which may not only be due to the 2D gration is fully explicit, using a second order accu-
steady-state assumptions in the RANS model, rate TVD Runge–Kutta scheme. The LES model
but may also be due to the RANS laminar flamelet consists of the filtered continuity, momentum,
approximation, where the unsteady coupling and energy equations together with the passive/re-
between the flow and the chemical reactions is active equations describing the flamelet models.
not accounted for. Moreover, the streamwise dis- Moreover, the mixed model, with scale similarity
tribution of h~vx i (not shown) along the geometrical and algebraic (Smagorinsky type) expressions, is
centerline of the combustor y/h = 4.16 is found to used for the subgrid turbulent viscosity, heat con-
be in good agreement with the experimental data, ductivity and diffusivity, as well as for the scalar
and in particular we find that the velocity drop dissipation rate and the variance of the mixture
behind the wedge is strong—much stronger than fraction. Two different flamelet models, using pas-
for the inert H2 injection case. This observation sive/reactive scalars and a global two-step and
is in good qualitative and quantitative agreement five-species reaction mechanism for H2-air are
with experimental data. used in this study.
Concerning the axial rms-velocity fluctuations In the scramjet engine model a mixing-con-
~vrms
x ¼ hð~vx  h~vx iÞ2 i1=2 , given in Fig. 4b, reason- trolled turbulent diffusion flame is anchored in the
able agreement is found between the present wake of the wedge-like strut, thus acting as a flame-
LES and the experimental data. The amplitude holder, whereas combustion is initiated in the shear
and the width are reproduced within a reasonable layers leaving the edges of the wedge. The combus-
accuracy, and the defect that is observed at the tion zone can be divided into an induction zone,
centerline is also found in the present LES. from the wedge base to the section where the diffu-
In Fig. 4c cross-sectional profiles of the time-av- sion flame starts to dominate, a transitional zone,
eraged temperature, hT~ i, are given at x/h = 13.00, where large-scale coherent structures are developed
20.83, 22.16, and 34.50 along with experimental together with convective mixing, and finally a tur-
data when available. At x/h = 13.00 we find that bulent (diffusion) flame zone.
both LES overpredict hT~ i in the shear layers, The pressure in the combustor is almost con-
whereas at the remaining cross-sections good stant, which is contrary to the non-reacting cases
agreement between the LES and the measurement also being investigated. Qualitative and quantita-
data is found although the statistical sampling is tive comparisons were systematically made with
marginal. It is interesting to observe the spatial evo- experimental data for both velocity and tempera-
lution of the temperature field with downstream tures. These comparisons were supplemented with
distance—the first cross-section is within the induc- comparisons of schlieren, OH-LIF (heat release)
tion zone, with mixing-controlled combustion, the and PIV (velocity) fields (not shown). In summa-
second is most likely within the transitional zone, ry, we find that the LES computations are capable
with maximum exothermicity, whereas the third is of predicting the flow reasonably well although
within the highly turbulent combustion zone. Fur- some discrepancies concerning the volumetric
thermore, we notice that the turbulent wall-bound- expansion and the time-averaged velocity are
ary layer increases the temperature by almost 200 K found. These deviations are however shared with
due to viscous heating—an effect that influences the other LES and RANS predictions.
thermal load.

Acknowledgments
6. Concluding remarks
DLR Institute of Space Propulsion, Lampolds-
In the present work LES has been used to
hausen, Germany, is acknowledged for providing
investigate mixing and combustion in a scramjet
the experimental data. Mr. H.G. Weller is
model under realistic operating conditions. The
acknowledged for the development of the C++
configuration used is similar to the laboratory
class library FOAM (Field Operation And
scramjet engine developed and experimentally
Manipulation), version 1.9.2 b, used here. The
investigated at DLR in Germany. The scramjet
authors also acknowledge Drs. T. Carlsson,
engine consists of a rectangular, one-sided diver-
K. Hannemann, and M. Oschwald for valuable
gent channel with a wedge-shaped flameholder
discussions as well as Dr. N. Wikström for
at the base of which hydrogen can be injected
fexcellent work on the computational meshes.
through an array of holes. In the computational
model periodicity is imposed in the spanwise
direction (with a span encompassing three holes)
to reduce the overall computational cost. References
The numerical method is based on an unstruc-
tured finite volume discretization, using TVD [1] T. Poinsot, D. Veynante, Theoretical and Numerical
reconstruction of the convective fluxes and central Combustion, R.T. Edwards, Philadelphia, 2001.
2504 M. Berglund, C. Fureby / Proceedings of the Combustion Institute 31 (2007) 2497–2504

[2] S. Menon, in: R. Friedrich, W. Rodi (Eds.), [14] K.N.C. Bray, Proc. Combust. Inst. 26 (1996) 1–26.
Advances in LES of Complex Flows, Kluwer, [15] R.C. Rogers, W. Chinitz, AIAA Paper, 1982, pp.
Dordrecht, 2002, p. 329. 82–0112.
[3] J. Janicka, A. Sadiki, Proc. Combust. Inst. 30 (2005) [16] E.S. Oran, J.P. Boris, Numerical Simulation of
537. Reactive Flow, Elsevier, New York, 1987.
[4] W.P. Jones, in: P.A. Libby, F.A. Williams (Eds.), [17] G. Erlebacher, M.Y. Hussaini, C.G. Speziale, T.A.
Turbulent Reacting Flows, Academic Press, London, Zang, ICASE Report No 90-76, NASA Langley,
1994, p. 309. 1990.
[5] S.B. Pope, Proc. Combust. Inst. 23 (1990) 591. [18] F. Gao, CTR Annual Research Briefs, Center for
[6] M. Oschwald, R. Guerra, W. Waidmann, in: Int Turbulence Research, 1993, p. 187.
Symp on Special Topics in Chem Prop, May 10–14, [19] W.P. Jones, P. Musonge, Phys. Fluids 31 (1988)
Scheveningen, NL, 1993, pp. 498–503. 3589–3604.
[7] W. Waidmann, F. Alff, U. Brummund, M. Böhm, [20] C. Duwig, C. Fureby, LES of unsteady stratified
W. Clauss, M. Oschwald, in: DGLR Jahrestagung, combustion, (in preparation).
Erlangen, 1994, pp. 629–638. [21] C. Fureby, Proc. Combust. Inst. 30 (2005) 593.
[8] W. Waidmann, U. Brummund, J. Nuding, in: 8th [22] C. Meneveau, T. Poinsot, Combust. Flame 86 (1991)
Int Symp on Transp Phenom in Comb, July 16–20, 311–332.
San Francisco, USA, 1995, pp. 1473–1484. [23] G.L. North, D.A. Santavicca, Combust. Sci. Tech-
[9] W. Waidmann, F. Alff, U. Brummund, M. Böhm, nol. 72 (1990) 215–232.
W. Clauss, M. Oschwald, Space Technol. 15 (1995) [24] S. Gottlieb, C.-W. Shu, Math. Comput. 67 (1998)
421–429. 73–85.
[10] M. Oevermann, Aerosp. Sci. Tech. 4 (2000) 463– [25] D. Drikakis, C. Fureby, F.F. Grinstein, M. Lief-
480. vendahl, In: F.F. Grinstein et al. (Eds.) Implicit
[11] F. Génin, B. Chernyavsky, S. Menon, AIAA Paper, Large Eddy Simulation: Computing Turbulent Fluid
2003, pp. 2003–7035. Dynamics, Cambridge University Press, (in press).
[12] C. Fureby, M. Berglund, in: E. Lamballais et al. [26] C. Fureby, N. Alin, N. Wikström, S. Menon, N.
(Eds.) Direct and Large Eddy Simulation VI, Svanstedt, L. Persson, AIAA J. 42 (2004) 457–468.
Springer, (in press). [27] M. Berglund, C. Fureby, LES of Flow, Mixing, and
[13] F.F. Grinstein, K.K. Kailasanath, Combust. Flame Combustion in a Supersonic Combustor, (in
100 (1995) 2–10. preparation).

Comments

Nickolay Smirkov, Moscow MV Lowouosov State d


University, Russia. The flamelet model you use
seems to be quite consistent with the LES model Jeong-Zeol Choi, Pusan National University, Korea.
in terms of basic assumptions. On the other hand, You mentioned the grid convergence by using velocity
the model accounting for the final rate of chemical profiles in an averaged field. But for turbulence combus-
reactions should take into account some subgrid tion it is considered that grid convergence should be con-
modeling of temperature fluctuations rather than firmed for turbulent characteristics such as turbulent
the mean values for temperatures due to the fact intensity and frequency to correctly account for the tur-
of strong exponential dependence of reaction rate bulence mixing that governs the physics. Would you
on temperature. clarify the level of grid resolution in your computation
with respect to the turbulence characteristics?
Reply. In this study, we are using two flamelet
models; the first is based on a transport equation for Reply. Two grids with 3 and 6 million grid points have
a passive scalar and an algebraic equation for its var- been used for this study and we have presented compari-
iance and the second is based on transport equations sons of the time-averaged axial velocity component, the
for a passive scalar and a reactive scalar. The thermo- temperature profiles, as well as the rms-fluctuations for
chemical state, in terms of species mass fractions, is the axial velocity component with good or reasonable
entirely determined by the flamelet libraries which in agreement with experimental data. Taking into account
turn are parameterized using the passive and reactive the allowed length of the paper and the information avail-
scalars, respectively. This approach entirely avoids able from the experiments, more detailed comparisons are
evaluating the Arrhenius rate expressions which indeed not possible. However, comparing statistical data be-
suffer from exponential temperature dependence. The tween LES predictions performed on different grids sug-
finite rate chemistry effects enter into the two-equation gests that the statistics of both the first and the second
flamelet model by means of the laminar flame speed order statistical moments are reasonably well captured
and its dependence on strain rate and the passive sca- on the coarse mesh. Some improvement can be observed
lar, i.e., the mixture fraction. for the fine grid in the shear layers in the transition region.

You might also like