A Comparative Study of Sandia Ame Series (D-F) Using Sparse-Lagrangian MMC Modelling

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Available online at www.sciencedirect.

com
Proceedings
of the
Combustion
Institute
Proceedings of the Combustion Institute 34 (2013) 1325–1332
www.elsevier.com/locate/proci

A comparative study of Sandia flame series (D–F)


using sparse-Lagrangian MMC modelling
Y. Ge a,⇑, M.J. Cleary a,b, A.Y. Klimenko a
a
The University of Queensland, School of Mechanical and Mining Engineering, QLD 4072, Australia
b
The University of Sydney, School of Aerospace, Mechanical and Mechatronic Engineering, NSW 2006, Australia

Available online 4 July 2012

Abstract

Sparse-Lagrangian Multiple Mapping Conditioning (MMC) simulations are performed on a turbulent


piloted methane jet flame series with increasing levels of local extinction (Sandia flames D–F). Subfilter
conditional scalar dissipation is closed by the generalised MMC mixing model which enforces localisation
in an extended space comprising of physical location and a reference mixture fraction; the latter is the fil-
tered field from an LES simulation. A new fractal/gradient model controlling localisation in sparse simu-
lations has been developed recently, and here it is tested on real flame conditions. Using one single
localisation constant fm, simulation results with as few as 1 Lagrangian particle per 27 Eulerian cells show
a good match to the measurements; particle number density effects are also explored.
Ó 2012 The Combustion Institute. Published by Elsevier Inc. All rights reserved.

Keywords: Filtered density function; Sparse-Lagrangian simulation; Multiple Mapping Conditioning; Fractal/gradient
model

1. Introduction function (FDF) [1] and LES/PDF [2] methods


provide a probabilistic closure of the subfilter
Strong turbulence-chemistry interactions such scales while preserving the exact chemical source
as found in turbulent combustion with local term. Practical FDF applications usually employ
extinction and re-ignition are the subject of much a hybrid implementation where the FDF is closed
research. Modelling is made difficult by the non- at the joint scalar level and simulated stochasti-
linearity of reaction rates and the small scales at cally using Pope particles (Lagrangian particles
which the interactions occur. This has implica- with properties and a mixing operation [3]), while
tions for large eddy simulation (LES) since the exploiting a well resolved turbulent velocity that is
processes which occur at the subfilter scales and simulated by an Eulerian LES. One evident fact of
which have to be modelled, are often more impor- this hybrid implementation is that the computa-
tant in determining flame structure (especially tional cost is linked with the number density of
pollutant formation) than those processes that particles. A typical FDF simulation uses 20–50
occur at the resolved scales. Filtered density particles per LES grid cell over a flow domain
encompassing a few million LES cells [4]. The cost
of simulating engineering-scale applications using
⇑ Corresponding author. Address: Room 310, Build-
such intensive-FDF methods is very high.
ing 45, The University of Queensland, QLD 4067, To make the FDF method accessible to engi-
Australia.
E-mail address: y.ge@uq.edu.au (Y. Ge).
neering scale problems, the sparse-Lagrangian

1540-7489/$ - see front matter Ó 2012 The Combustion Institute. Published by Elsevier Inc. All rights reserved.
http://dx.doi.org/10.1016/j.proci.2012.06.059
1326 Y. Ge et al. / Proceedings of the Combustion Institute 34 (2013) 1325–1332

FDF method, which uses significantly fewer parti- variance can be obtained by one set of localisation
cles than LES grid cells, has been developed in parameters with relative insensitivity to particle
recent years [5] and tested against laboratory number density and reaction zone width [10]. This
flames with moderate to high levels of local extinc- now needs to be tested for a wide range of real
tion [6,7]. The success of the sparse-Lagrangian flames with finite-rate chemistry. The aims of the
method relies on the MMC mixing model which present paper are to examine universality of the
enforces mixing localisation even under the sparse model parameters for a series of partially pre-
conditions. A particle interaction micromixing mixed methane flames with varying levels of local
model is used where particles are selected to form extinction (Sandia flames D–F) [13]. The major
mixing pairs which are close in an extended space details of the model with emphasis on the frac-
of physical location and reference variables. For tal/gradient localisation model are presented in
the diffusion flames cases simulated so far the the next section, followed by a description of the
reference variable is given by the LES filtered mix- simulation cases in Section 3. Results addressing
ture fraction. This choice is made because localisa- the aims of the paper appear in Section 4 and con-
tion in mixture fraction implies a localisation in clusions are drawn in Section 5.
the composition space. For non-local mixing
models like IEM [8], which mixes with a local fil-
tered mean, and Curl’s model [9], which mixes 2. The model
randomly selected particle pairs from within the
same grid cell, the extent of localisation in compo- The computational model adopted in this com-
sition space is determined implicitly by the size of parative study is the same as that previously devel-
the filter and the number of particles that are used; oped for the separate Sandia flame D [6] and E
the smaller the filter and the greater the number of studies [7] and which is presented in detail in
particles the better the localisation. Such localisa- [10]. A brief overview of the hybrid LES/sparse-
tion by numerical intensification is limited by Lagrangian MMC model is given here but readers
computational resources. If non-local mixing are directed to the previous publications for com-
models were applied in sparse-Lagrangian FDF plete details. The model involves an Eulerian LES
simulations, the distance between mixing particles with a dynamic Smagorinksy turbulent viscosity
in composition space would be large. It would model for the filtered velocity and reference mix-
result in non-physical mixing of particles across ture fraction fields and a stochastic Lagrangian
the stoichiometric reaction zone [10] and would model for the reactive scalars. Density feedback
likely lead to failure of the model. In contrast, from the Lagrangian to the Eulerian scheme uses
localised mixing models, such as MMC, require a conditionally averaged form of the equivalent
that mixing particles in the vicinity of stoichiome- enthalpy method originally developed by Mura-
try are close in the composition space and thus doglu et al. [14] and used in intensive-Lagrangian
they give better predictions. FDF simulations by Raman and Pitsch [4]. The
Sparse-Lagrangian MMC has been applied in adapted density feed back method matches the
separate Sandia flame D [6] and E studies [7] at conditional means of the Eulerian and Lagrangian
the two most recent past symposia. That model- submodels. It is fully documented and tested
ling relied on a selectable localisation parameter, against flames with real and ideal chemistry in
k. Problematically k is not related to the physical Refs. [7,10].
quantities of the flow and does not explicitly The subfilter conditional scalar dissipation in
reveal the localisation structure of the mixing the FDF is closed through a particle mixing
model. Furthermore its value, which was tuned model. Mixing pairs are selected by an approxi-
by a priori simulations, varies with the number mate minimisation of the normalised square dis-
of particles used. A practical combustion model tance in the extended space of physical location
should have a universal set of parameters, at least and reference mixture fraction, ðx; f~ Þ-space, [7]:
for flames within a series; this has been demon- 2 0pffiffiffi p;q 12 !2 3
strated by Xu and Pope [11] and Jones and Prasad X3 3d ðx Þ d p;q
ðf Þ
[12] for the Sandia series. To this end a new frac- d^ðp;qÞ ¼ 4 @
2 j
A þ 5: ð1Þ
j¼1
rm fm
tal/gradient scale model for localisation of sparse-
Lagrangian MMC was introduced [10]. The
In (1) asterisks denote stochastic quantities as-
model parameters are explicitly linked to the
signed to or evaluated at the particles and
localisation structure allowing the distance p q
d p;q
ð:Þ ¼ jð:Þ  ð:Þ j. The parameters rm and fm are
between mixing particles in the reference mixture
characteristic scales in the physical and reference
fraction space to be directly controlled. This is
spaces, respectively. In the previous publications
important, especially for local extinction and re-
[7] they were set as the geometric scales of the
ignition, because that distance is the most impor-
flow; rm was set as the nozzle radius and fm = 1/
tant parameter controlling conditional variance.
k, where k was the localisation control parameter.
Simulations of an idealised flame with infinite-rate
In this paper a new fractal/gradient model [10] is
chemistry show that predictions of conditional
Y. Ge et al. / Proceedings of the Combustion Institute 34 (2013) 1325–1332 1327

to be used. The model mathematically links the defined point where the scalar gradient d f~ =dn is
scales rm and fm by considering isoscalar contours defined by the boundary conditions and is the
in a turbulent field as fractal surfaces. A mixing same for all three flames in the Sandia series.
particle pair is selected so that both particles lie
close to the same fractal surface such that they
are separated by a small distance, fm, in reference 3. Simulation details
mixture fraction space. By equating the volume
subtended by the particle pair and the volume of The comparative study using sparse-Lagrang-
fluid represented by each particle (as defined by ian MMC is performed for the well documented
the particle number density) it is found that [10] Sandia flame series D–F [13]. This series has
!1=D increasing jet velocities and exhibits increasing
d f~ D3L 1 levels of local extinction with flame F being close
r m ¼ cm : ð2Þ to blow-off. The burner has a central fuel nozzle of
dn D2D
E
fm
diameter d = 7.2 mm surrounded by a pilot that
Here cm is a constant, d f~ =dn is the gradient of the extends to a diameter of 18.2 mm. The jet fuel is
reference mixture fraction normal to the fractal 25% CH4 and 75% air by volume and has a stoi-
surface, DL is the nominal distance between parti- chiometric mixture fraction of Zst = 0.351. The
cles (linked to the particle number density), DE is pilot burns a lean premixture with the same nom-
the LES filter width and D is the fractal dimen- inal enthalpy and equilibrium composition as
sion. Fractal properties in turbulence are dis- methane/air at an equivalence ratio of 0.77. The
cussed in a number of publications [15,16]. The bulk jet velocities are 49.9, 74.4 and 99.2 m/s for
computational fractal dimension that achieves flames D, E and F, respectively. The correspond-
good control of mixing distances over a range of ing pilot inlet velocities are 11.4, 17.1 and
sparse conditions [10] seems to match the experi- 22.8 m/s.
mentally observed fractal dimension of 2.36 [15]. The hybrid Eulerian LES/sparse-Lagrangian
Comparison between predicted (from Eq. (2)) MMC model is implemented numerically within
and numerically observed inter-particle distances the Flowsi LES code [18] as documented in earlier
from our previous simulations indicated that publications [5,6]. The cylindrical simulation
cm = 0.5. Note that cm is not related to the micro- domain extends axially to Laxial = 0.25 m (35 jet
mixing timescale as, say, C/ is in RANS-PDF diameters) and radially to Lrad = 0.25 m. The
models. In sparse simulations in LES the mixing mesh used consists of 512  55  32 in the axial,
timescale is determined by the scaling laws of radial and azimuthal directions, respectively. The
the inertial interval. That modelling is docu- smallest finite volume cells at the axis are
mented in detail elsewhere [7]. 0.5 mm  0.5 mm  p/32 rad. In the sparse-
We now see that the scales rm and fm are linked Lagrangian MMC we use 1 particle per 27 Euleri-
to each other through Eq. (2) rather than arbi- an LES cells (1L/27E). The nominal distance
trarily to the geometric scales of the flow and k between particles at the jet axis is approximately
as was previously the case [7]. fm represents a char- DL = 1.6 mm. After a steady-state has been
acteristic distance in mixture fraction space reached, stationary statistics are accumulated over
between mixing particles. rm is the characteristic eight characteristic domain flow through times.
physical distance between mixing particles and as Chemical source terms are obtained from a
such is viewed as a characteristic Lagrangian filter detailed kinetics scheme (GRI-3.0) [19] containing
width. fm is selected by the user; a very small value 34 species and 219 reactions (NOx excluded). To
would suppress conditional fluctuations and pro- test sensitivity of the model to the Lagrangian fil-
duce a model with similar properties to first-order ter width (characterised by rm), one additional
conditional moment closure [17] whereas larger simulation is run for flame F using 1 particle per
values of fm produce conditional fluctuations. 8 LES cells (1L/8E). We chose this case because,
The optimal value of fm will depend on the flame being the most difficult of the three flames, predic-
regime (i.e. how well the fluctuations of the tions are the most sensitive to variations in the
mixture fraction define the turbulence-chemistry model.
interactions). An analysis of the distances between
mixing particles of our past simulations [7] indi-
cates that fm = 0.03 should be used for the partially 4. Results
premixed methane flames in the Sandia series.
As previously described [6] the numerical algo- To best demonstrate the extinction and re-igni-
rithm for mixing pair selection approximately tion phenomena, predictions are analysed at x/
minimises d^2ðp;qÞ using a kd-tree method which d = 7.5 and 15 where strong extinction is evident
requires global values for fm and rm. Therefore and at x/d = 30 where re-ignition is nearly com-
we calculate a global value of rm based on the plete. Unless otherwise noted predictions are for
quantities in Eq. (2) at a characteristic point in the 1L/27E simulations. Some figures also contain
the flow. We use the nozzle exit since this is a well
1328 Y. Ge et al. / Proceedings of the Combustion Institute 34 (2013) 1325–1332

1L/8E predictions; these will be discussed towards


the end of the results section.
The MMC model, unlike conventional mixing
models such as IEM and Curl’s, allows for quite
direct control of the conditional variance. This is
done by controlling the distance between mixing
particles in the reference mixture fraction space.
Figure 1 shows the numerically observed mean
distance between mixing particles in that space,
hd f  i, for all three flames. Note that the relation-
ship between hd x i and hd f  i is the same as the
mathematical relationship between rm and fm
given by Eq. (2); it is just that the former are the
actual observed mixing distances in physical and
reference mixture fraction spaces at any location
in the flow, while the latter are the characteristic
values used in Eq. (1). It can be seen from the fig-
ure that hd f  i peaks around the shear layer where
the scalar gradient, d f~ =dn, is largest. For 1L/27E
simulations hd f  i for all three flames are in good
agreement with each other. Close to the nozzle
exit location, x/d = 0.5, the maximum value of
hd f  i is reasonably well controlled to 0.03 which Fig. 2. Radial profiles of steady-state mean and rms for
is the preset characteristic value for the model mixture fraction. Open symbols, experimental data;
parameter, fm. It needs to be pointed out that lines, 1L/27E simulation results. From top to bottom
hd f  i varies by a lot at different locations in the are flame D, flame E and flame F results.
flow according the magnitude of the scalar gradi-
ent. The dip at r/d = 1 at the nozzle exit coincides
with the pilot region where the scalar gradient is and lower density than in the experiments, yield-
near zero. ing slightly slower decay of the jet core.
Figure 2 shows radial profiles of the steady- Predicted and experimental scatter plots of
state mean and rms of the mixture fraction. Note temperature versus mixture fraction are shown
this is the actual mixture fraction determined from in Figs. 3–5 and the corresponding conditional
the sparse-Lagrangian MMC 1L/27E simulation means are shown later in Fig. 6. With increasing
and not the reference mixture fraction given by jet velocity, the flame series exhibit an increasing
the Eulerian LES. An excellent agreement level of extinction, most evident in flame E and
between the simulations and experiments are F at x=d ¼ 7:5 and 15. Predictions for flames D
achieved for both flames D and E at all axial loca- and E are in very good agreement with the exper-
tions. Flame F is the most difficult of the cases to imental data. There are slightly fewer conditional
model correctly due to its being close to blow-off. fluctuations evident in the predicted scatter plots
The mixture fraction predictions for the flame F for flame D but this flame has very low condi-
case appear to be less diffusive than the experi- tional variance anyway and the conditional means
ment; this might be attributed to a slight under are in excellent agreement at all three axial loca-
prediction of extinction for flame F (see discussion tions. As a close to blow-off flame, the bimodal
below) which in turn implies higher temperatures nature of flame F is clearly captured in the pre-
dicted scatter plot at x=d ¼ 15; there is an upper
flamelet band and a lower close to extinction band
which peaks at about 1000 K. For the flame F
simulation, the scatter plot at x=d ¼ 30 predicts
an earlier re-ignition relative to the experiments.
Similar discrepancy has also been reported by oth-
ers [12]. In our simulations great care was required
to avoid the occurrence of isolated turbulent
events where the temperature of a small number
of particles departed a great distance from the
mean leading to global extinction. These rare
events are particularly important for very sparse
Fig. 1. Radial profiles of steady-state mean distance in simulations where each particle represents a sig-
mixture fraction space between mixing particles. Solid nificant mass of fluid. We tried increasing fm to
line, flame D; dashed line, flame E; doted line, flame F; 0.035 to better capture the level of local extinction
star, flame F with 1L/8E. in flame F. The additional simulations were tried
Y. Ge et al. / Proceedings of the Combustion Institute 34 (2013) 1325–1332 1329

Fig. 3. Scatter plots of temperature versus mixture Fig. 5. Scatter plots of temperature versus mixture
fraction for flame D. Top, x/d = 7.5; middle, x/d = 15; fraction for flame F. Top, x/d = 7.5; middle, x/d = 15;
bottom, x/d = 30. bottom, x/d = 30.

Fig. 6. Steady-state conditional mean temperature ver-


sus mixture fraction. From top to bottom are flame D,
flame E and flame F. Open symbols, experimental data;
Fig. 4. Scatter plots of temperature versus mixture solid lines, 1L/27E simulation results; dashed-lines,
fraction for flame E. Top, x/d = 7.5; middle, x/d = 15; results for flame F with 1L/8E.
bottom, x/d = 30.
turbulent events that in some cases would occur
for flames D and E also. While this lead to more a few domain flow through times into the simula-
pleasing scatter plots for flame D (but no signifi- tion after the steady state had (apparently)
cant change in conditional mean profiles) it already been reached. Another issue is that we
caused flame F to globally extinguish due to rare use fm as a global value. The scatter plots for flame
1330 Y. Ge et al. / Proceedings of the Combustion Institute 34 (2013) 1325–1332

Fig. 7. Steady-state conditional mean CO2 versus mix- Fig. 9. Radial profiles of steady-state mean and rms of
ture fraction. From top to bottom are flame D, flame E CO2. From top to bottom are flame D, flame E and
and flame F. Open symbols, experimental data; solid flame F. Diamonds, experimental mean; triangles,
lines, 1L/27E simulation results; dashed-lines, results for experimental rms; solid lines, 1L/27E simulation results;
flame F with 1L/8E. dashed-lines, results for flame F with 1L/8E.

Fig. 8. Steady-state conditional mean CO versus mix-


ture fraction. From top to bottom are flame D, flame E Fig. 10. Radial profiles of steady-state mean and rms of
and flame F. Open symbols, experimental data; solid CO. From top to bottom are flame D, flame E and flame
lines, 1L/27E simulation results; dashed-lines, results for F. Diamonds, experimental mean; triangles, experimen-
tal rms; solid lines, 1L/27E simulation results; dashed-
flame F with 1L/8E.
lines, results for flame F with 1L/8E.

F reveal reasonable agreement with experiments parameters is possible in principle but has not
at x/d = 7.5 and 15 but the less satisfactory result been tested here.
at x/d = 30 suggests a larger value of fm may be The conditional means of the reactive scalars
required there. Having fm determined by local CO2 and CO are shown in Figs. 7 and 8. Again
parameters rather than global characteristic a very good agreement between simulation and
Y. Ge et al. / Proceedings of the Combustion Institute 34 (2013) 1325–1332 1331

experiment results are achieved for flames D and to the physical distance between mixing particles
E; there is some slight over-prediction of rich side (characterised by rm). Beyond that threshold level
CO in flame E. For flame F only the lean side is of local extinction (i.e. the extent seen at x/
well matched, however, in line with the tempera- d = 15) it is not enough just to control the mixing
ture results discussed above the stoichiometric distance in mixture fraction space by the single glo-
and rich side values are over-predicted. The bal parameter fm; other quantities start play a sig-
unconditional mean and rms profiles of CO2 and nificant role e.g. physical distance and/or a
CO are shown in Figs. 9 and 10. In general there progress variable. This is perhaps not too surpris-
is very good agreement between the predicted and ing since heavy local extinction leads to premixing
measured results, although there is an over-pre- and less of a correlation between fluctuations of
diction of the intermediate CO in flame F which mixture fraction and reactive species.
corresponds to the conditional average results.
The above comparative results are from simula-
tions with 1 particle per 27 Eulerian cells (1L/27E). 5. Conclusion
Most importantly the results demonstrate that
with a single value of fm the model correctly pre- The Sandia flame series D–F have been simu-
dicts the trend of increasing local extinction with lated using the sparse-Lagrangian MMC method.
increasing jet velocity. An advantage of sparse The detailed localisation structure in the MMC
methods is the ability to control the number of mixing model is controlled according to the phys-
particles and to balance computational cost and ically-based fractal/gradient scaling model. The
level of detail in the predictions (e.g. stationary sta- analysis shows that the fractal/gradient model
tistics can be achieved with very few particles while maintains the same distance in reference mixture
larger numbers of particles are needed to accu- fraction space between mixing particles in all three
rately observe local and instantaneous filtered cases. This model also allows particle number den-
means [10]). The new fractal/gradient localisation sity to be adjusted. When fm is set to 0.03 the pre-
model attempts to allow the same value of fm to dicted conditional and unconditional reactive
be used for simulations with different numbers of scalars are in very good agreement with experi-
particles. The aim is to keep fm the same regardless mental data for flames D and E at all locations
of the number of particles used while rm varies in a in the flow. For flame F, which is close to blow-
known way according to Eq. (2). Here this is tested off, results are qualitatively correct, the bimodal
by an additional 1L/8E simulation for flame F. nature of the extinction is captured, but there is
Since particle number density is increased rm an early prediction of re-ignition. Importantly
becomes smaller. This can be viewed as a reduction the increased level of extinction arising from the
in the characteristic Lagrangian filter width. Since increasing jet velocity is successfully captured by
rm is changed, the 1L/8E simulation is not a the model with a single value of fm. The sensitivity
numerical convergence test but rather a model sen- of the model to the physical distance between mix-
sitivity test. Note that numerical convergence in ing particles has been analysed with a more inten-
sparse methods while keeping the Lagrangian filter sive 1L/8E simulation for the most difficult case
width constant is discussed theoretically in Ref. [3]. (flame F). The fractal/gradient model effectively
Figure 1 (discussed previously) shows that hd f  i for controls the distance in reference mixture fraction
1L/8E is very similar to that for the 1L/27E simu- space between mixing particles to be the same as
lation though for 1L/8E there are slightly smaller the more sparse simulation. Close to the nozzle
peak values. This very minor discrepancy is prob- and in the far downstream predictions are insensi-
ably caused by numerical noise in the particle tive to the particle number density. However, the
number control algorithm [6]; in very sparse simu- reactive scalar predictions at the location with
lations the relative fluctuations of particle number strongest local extinction are sensitive. Clearly
density are larger. The conditional and uncondi- where there is heavy extinction localness of mixing
tional means predicted by the 1L/8E simulation in mixture fraction space does not ensure full
are shown alongside the 1L/27E predictions in localness in composition space. This problem
Figs. 6–10. We see that overall there is good simi- may be overcome by local rather than global con-
larity between the results although there are some trol parameters, fm and rm, or the incorporation of
local differences. At x/d = 7.5 and 30 the agree- an additional reference variable such as sensible
ment is excellent for all quantities but there is dis- enthalpy as has been done in the past for the
crepancy at x/d = 15 which is the point in the deterministic form of MMC.
flame with maximum local extinction. These
results indicate that up to a certain level of local
extinction (i.e. the extent seen at x/d = 7.5) the dis- References
tance in mixture fraction space between mixing
particles (characterised by fm) is the most impor- [1] F.A. Jaberi, P.J. Colucci, S. James, P. Givi, S.B.
tant model parameter for controlling conditional Pope, J. Fluid Mech. 401 (1999) 85–121.
variance and predictions are relatively insensitive [2] S.B. POPE, J. Fluid Mech. 652 (2010) 139–169.
1332 Y. Ge et al. / Proceedings of the Combustion Institute 34 (2013) 1325–1332

[3] A. Klimenko, M. Cleary, Flow Turbul. Combust. 85 [12] W. Jones, V. Prasad, Combust. Flame 157 (2010)
(2010) 567–591. 1621–1636.
[4] V. Raman, H. Pitsch, Proc. Combust. Inst. 31 (2007) [13] R.S. Barlow, J.H. Frank, Sympos. (Int.) Combust.
1711–1719. 27 (1998) 1087–1095.
[5] M.J. Cleary, A.Y. Klimenko, Flow Turbul. Com- [14] M. Muradoglu, S.B. Pope, D.A. Caughey, J.
bust. 82 (2009) 477–491. Comput. Phys. 172 (2001) 841–878.
[6] M.J. Cleary, A.Y. Klimenko, J. Janicka, M. Pfitz- [15] K.R. Sreenivasan, Annu. Rev. Fluid Mech. 23
ner, Proc. Combust. Inst. 32 (2009) 1499–1507. (1991), 539-+.
[7] Y. Ge, M.J. Cleary, A.Y. Klimenko, Proc. Com- [16] V.R. Kuznetsov, V.A. Sabelnikov, Turbulence and
bust. Inst. 33 (2011) 1401–1409. Combustion, Hemisphere, New York, 1990.
[8] C. Dopazo, Phys. Fluids 18 (1975) 397–404. [17] A.Y. Klimenko, R.W. Bilger, Prog. Energy Com-
[9] R.L. Curl, AIChE J. 9 (1963) 175–181. bust. Sci. 25 (1999) 595–687.
[10] M.J. Cleary, A.Y. Klimenko, Phys. Fluids 23 (2011) [18] A. Kempf, A. Sadiki, J. Janicka, Proc. Combust.
115102. Inst. 29 (2002) 1979–1985.
[11] J. Xu, S. Pope, Combust. Flame 123 (2000) 281–307. [19] Available at http://www.me.berkeley.edu/gri_mech/.

You might also like