DNS Study of The Optimal Heat Release Rate Marker in Premixed Methane Flames

You might also like

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

JID: PROCI

ARTICLE IN PRESS [m;August 4, 2018;5:21]

Available online at www.sciencedirect.com

Proceedings of the Combustion Institute 000 (2018) 1–9


www.elsevier.com/locate/proci

DNS study of the optimal heat release rate marker in


premixed methane flames
Cheng Chi, Gábor Janiga, Katharina Zähringer, Dominique Thévenin∗
Laboratory of Fluid Dynamics and Technical Flows (LSS/ISUT), University of Magdeburg “Otto von Guericke”,
Universitätsplatz 2, Magdeburg, D-39106, Germany
Received 30 November 2017; accepted 21 July 2018
Available online xxx

Abstract

The objective of this study is to identify a chemical marker combination that can be measured by Laser-
Induced Fluorescence and used for predicting the heat release rate distribution in premixed methane-air
flames at various equivalence ratios. Direct Numerical Simulations of turbulent methane flames at different
equivalence ratios have been performed to obtain an indirect but accurate estimation of heat release based
on all possible chemical markers. Two detailed kinetic mechanisms have been considered to check that a ro-
bust estimation is obtained. The simulation results are post-processed and analyzed by a pixel-to-pixel image
comparison. At the difference of other studies, the normalized distribution of heat release and the reconstruc-
tion based on chemical markers are compared directly, so that it is possible to investigate as well geometric
flame parameters, like flame thickness, for specific regions of heat release. Optimal marker combinations for
measurable species have been combined with corresponding exponents. Finally, for methane flames with an
equivalence ratio ranging from 0.6 to 1.4, it is found that the marker combination ([OH]1.07 × [CH2 O]1.17 )
(where [ ] denotes the normalized concentration) leads to a very high reconstruction quality for heat release,
particularly so at high levels of heat release. Alternatively, ([H]1.11 × [CH2 O]0.86 ) leads to an even better ap-
proximation of the whole heat release field. However, OH-based markers are easier to measure with a good
accuracy.
© 2018 The Combustion Institute. Published by Elsevier Inc. All rights reserved.

Keywords: DNS; Methane flames; Heat release rate; Reconstruction; LIF

1. Introduction technology. Heat release is also used for flame front


identification [1]. Thus, an accurate measurement
The spatial distribution of heat release is one of heat release in laminar and turbulent flames is
of the most important quantities in combustion of high importance both for fundamental research
and practical applications. However, a direct mea-
surement of heat release rate is not possible us-
∗ Corresponding author. ing current experimental techniques in turbulent
E-mail addresses: cheng.chi@ovgu.de, flames. Thus, many projects have considered over
thevenin@ovgu.de (D. Thévenin). the last decades indirect measurement techniques

https://doi.org/10.1016/j.proci.2018.07.095
1540-7489 © 2018 The Combustion Institute. Published by Elsevier Inc. All rights reserved.

Please cite this article as: C. Chi et al., DNS study of the optimal heat release rate
marker in premixed methane flames, Proceedings of the Combustion Institute (2018),
https://doi.org/10.1016/j.proci.2018.07.095
JID: PROCI
ARTICLE IN PRESS [m;August 4, 2018;5:21]

2 C. Chi et al. / Proceedings of the Combustion Institute 000 (2018) 1–9

that can be used to approximate the heat release impossible in turbulent flames using the current di-
profile based on measurable species or radicals dis- agnostic techniques.
tributions. In this study, a systematic evaluation of all pos-
The most widely used techniques to indi- sible HRR marker combinations is done for species
rectly track the heat release rate (HRR) are (1) measurable by LIF in methane-air flames. Direct
chemiluminescence, and (2) laser -induced fluores- Numerical Simulations (DNS) of turbulent flames
cence (LIF). The chemiluminescence from excited have been carried out for a variety of equivalence
molecules OH∗ and CH∗ has been reported as re- ratios. Building on top of previous studies for sim-
liable HRR indicator in [2–5]. Though chemilumi- pler fuels [20,21], a direct pixel-by-pixel compari-
nescence measurements are relatively simple, they son is proposed to identify optimal HRR markers
only deliver a signal integrated along the line of measurable by LIF for turbulent methane flames,
view. It is possible to characterize heat release by while minimizing the sensitivity toward the equiv-
chemiluminescence for simple laminar flames, as alence ratio. Different from the previous study on
extensively discussed, e.g., in [6]. However, such syngas flames [20], the current study investigates on
measurements are extremely challenging in turbu- methane flames, which are far more complex and
lent configurations, where computed tomographic interesting. This is also the first time that so many
approaches are necessary [7]. Additionally, chemi- different configurations have been considered in the
luminescence measurements of OH∗ and CH∗ were analysis, ensuring robustness of the predictions. A
found to be an unreliable HRR marker for local ex- more detailed analysis has been introduced, involv-
tinction by Najm et al. [8,9]. ing additionally joint probability density functions
As an alternative, HCO PLIF (Planar LIF) (PDFs). Most importantly, the current study is the
showed a strong correlation with local HRR. How- first time that a separate analysis is done for differ-
ever, due to the low HCO concentration, low flu- ent regions of heat release. In this manner, it is ob-
orescence quantum yield, and short excited-state served that a generally valid HRR marker can be
lifetime of HCO in flames, its measurement suf- used to track simultaneously the highly active re-
fers from a poor signal-to-noise ratio. To over- action zone and extinction events. Additionally, a
come this difficulty, Paul and Najm [9] proposed a feasibility study concerning the proposed marker is
pixel-by-pixel product of the OH and CH2 O PLIF discussed for LIF experiments, which has not been
images. In contrast to the HCO signal, the (OH done in our previous works [20,21].
× CH2 O) images show a good quality and rela-
tively intense signals in both laminar and turbulent
flames. and can be obtained rather easily. This inter- 2. Numerical simulations
esting method has, since then, been used and val-
idated by many different authors for a wealth of Direct Numerical Simulations of premixed tur-
applications [10–15]. Nikolaou and Swaminathan bulent flames have been performed using the
[16] later reexamined the validity of the reaction in-house low-Mach reacting flow solver DINO
rate of OH + CH2 O ⇔ HCO + H2 O as a re- [22,23]. Starting from an initially spherical reaction
liable heat release rate marker for lean methane- zone, a premixed flame propagates into a turbulent
air and preheated stoichiometric multicomponent environment.
mixtures (involving CO, H2 , H2 O, CO2 , and CH4 ). To check the robustness of the obtained pre-
For lean mixtures, it was found that the HRR dictions, two detailed reaction mechanisms for
correlates stronger with the rate of reaction OH methane flames have been employed: (1) GRI
+ CH2 O ⇔ HCO + H2 O at higher heat release. Mech 3.0 [24], and (2) the San Diego mechanism
For near-stoichiometric mixtures, the HRR corre- [25]. Both mechanisms have been widely used for
lates better with the reaction H + CH2 O ⇔ HCO methane combustion. Additionally, the reduced
+ H2 . The performance of (H × CH2 O) was mechanism of Smooke [26] has been employed
recently compared with the conventional marker in 3D (Case M in Table 1) to check the validity
(OH × CH2 O) [17]. It was found that both of the results obtained in two-dimensional DNS-
(H × CH2 O) and (OH × CH2 O) are suitable simulations (Cases A-I0 in Table 1). All chemi-
markers for methane-air flames with equivalence cal kinetics, thermodynamic, and transport coef-
ratio φ = 0.8–1.1, but that (H × CH2 O) is less ficients are systematically computed using Can-
sensitive to the equivalence ratio. Even more re- tera 1.8 [27].
cently, Wei et al. [18] studied the impact of equiva- A homogeneous isotropic turbulence field is ini-
lence ratio on HRR markers. They found that the tiated using a parallel turbulence generator based
correlations (H × CH2 O) and (OH × CH2 O) are on an Inverse Fast Fourier Transform relying on
sensitive to the mixture equivalence ratio, resulting the von Kármán spectrum with Pao correction [28].
in a large prediction uncertainty for specific con- The relevant turbulence parameters for each case
ditions, such as stratified flames. They proposed are summarized in Table 1. Here, φ is the mixture
(O × CH3 ) to be the best correlation for a wide equivalence ratio, sl is the laminar flame speed for
range of equivalence ratios; however, O-atom LIF the corresponding initial conditions. The Reynolds
is very challenging [19] and CH3 LIF is even number is Re = urms lint /ν r , where urms , lint and ν r

Please cite this article as: C. Chi et al., DNS study of the optimal heat release rate
marker in premixed methane flames, Proceedings of the Combustion Institute (2018),
https://doi.org/10.1016/j.proci.2018.07.095
JID: PROCI
ARTICLE IN PRESS [m;August 4, 2018;5:21]

C. Chi et al. / Proceedings of the Combustion Institute 000 (2018) 1–9 3

Table 1
Investigated DNS cases.
Case Mech φ sl (m/s) urms /sl lint /δ Re η (μm)
A GRI 3.0 0.6 1.52 3.51 21.6 92.9 49.2
B GRI 3.0 0.7 1.84 2.89 25.9 93.2 49.3
C GRI 3.0 0.8 2.10 2.52 29.5 91.5 49.5
D GRI 3.0 0.9 2.26 2.35 32.4 93.4 49.7
E GRI 3.0 1.0 2.33 2.28 31.4 90.1 49.5
F GRI 3.0 1.1 2.30 2.31 33.4 94.2 49.9
G GRI 3.0 1.2 2.16 2.47 31.0 94.0 49.7
H GRI 3.0 1.3 1.91 2.79 27.6 94.4 49.9
I GRI 3.0 1.4 1.59 3.35 22.0 90.0 49.5
J GRI 3.0 1.0 2.33 1.55 23.2 43.9 60.7
K GRI 3.0 1.0 2.33 3.19 44.4 173.4 41.5
A0 San Diego 0.6 1.54 3.45 21.9 91.9 49.1
B0 San Diego 0.7 1.91 2.79 26.2 92.8 49.3
C0 San Diego 0.8 2.13 2.50 29.7 92.5 49.3
D0 San Diego 0.9 2.27 2.35 33.0 93.6 49.6
E0 San Diego 1.0 2.35 2.26 32.3 93.1 49.5
F0 San Diego 1.1 2.32 2.29 33.6 94.2 49.9
G0 San Diego 1.2 2.16 2.47 31.3 93.7 49.5
H0 San Diego 1.3 1.89 2.82 26.5 91.4 49.2
I0 San Diego 1.4 1.57 3.39 22.6 90.4 49.4
M Smooke 1.0 2.32 2.26 166.8 461 32.9

are the fluctuating velocity, integral length scale From the fluorescence signal modeling in [31],
and kinematic viscosity on the reactant side, respec- the normalized distribution of the LIF signal in-
tively. The Zeldovich thickness is defined as δ = tensity [Sf ] reads
νr /sl , and η is the Kolmogorov length scale. These
[S f ] = f (T, p, χ p )[χabs ], (1)
flames span from thin reaction zones to corrugated-
flamelet regimes in the turbulent premixed combus- [χ abs ] is the normalized distribution of the molar
tion diagram of Peters [29]. concentration of the excited species, and f(T, p, χ p )
The computational domain has a side contains the combined effects of temperature T,
length L = 1 cm. The domain is discretized by pressure p, and mole fraction of perturbing species
1024 × 1024 grid points, resulting in a resolution χ p . In the present study, all flames are atmospheric
of 9.8 μm. This resolution is fine enough to resolve and the pressure is nearly constant within the flame,
both stiff radical species like CH2 O or H2 O2 , so that pressure-dependency is not an issue. The
and the Kolmogorov length scale. An initial hot net influence of the perturbing species correlates
kernel with a temperature Tr = 1900 K and a with the temperature and has an inverse power-
radius r = 1.5 mm is initialized at the center of the law dependence [17,32]. Therefore, this effect can
domain and used to ignite the surroundings, set at be lumped together with the temperature. The tem-
a temperature Tu = 800 K like in [16]. The initial perature dependence appears through the quench-
pressure is p = 1 atm. All the domain boundaries ing rate coefficient and the Boltzmann population.
are periodic. The simulations have been continued When selecting properly the excited transition line,
up to t = 1.5τt in order to reach a proper equi- the Boltzmann population is insensitive to temper-
librium between turbulence and chemistry [30], ature. For OH LIF, The quenching rate depends
where τt = lint /urms is the characteristic time scale only weakly on temperature [33]. The quenching
of the time-decaying turbulence. The simulation rate of H LIF can be assumed to be proportional
results at t = 1.5τt are used to identify optimal to T −1 [34], while CH2 O LIF is assumed to be pro-
HRR markers. portional to T −0.75 [17]. Regardless of the specific
temperature-dependence model, the overall effect
at the condition of two-photon overlap is given by
3. Identifying markers of heat release f(T),
[S f 1 ] × [S f 2 ] = f (T )[χ1 ] × [χ2 ]. (2)
As explained previously, it is necessary to ob-
tain HRR markers that are able to predict heat re- where [Sf1 ], [Sf2 ], [χ 1 ] and [χ 2 ] are the normalized
lease for a wide range of equivalence ratios. Species distribution of LIF signal intensity of species 1
measurable by LIF are considered to be H, OH, and species 2, and normalized concentration of
CH2 O, and HCO, like in [16]. As in previous studies species 1 and species 2, respectively. Knowing the
[20,21], species have been combined with exponents details of the experimental setup, it is usually not
in order to improve the accuracy of the prediction. difficult to obtain the correct model for f(T) [35].

Please cite this article as: C. Chi et al., DNS study of the optimal heat release rate
marker in premixed methane flames, Proceedings of the Combustion Institute (2018),
https://doi.org/10.1016/j.proci.2018.07.095
JID: PROCI
ARTICLE IN PRESS [m;August 4, 2018;5:21]

4 C. Chi et al. / Proceedings of the Combustion Institute 000 (2018) 1–9

Since the model depends on the setup, a specific


modeling for f(T) will not be discussed in the
present study, relying purely on numerical simula-
tions. For two species LIF, the two camera views
(LIF signals) are overlapped on a pixel-by-pixel ba-
sis to evaluate HRR. As the distribution of the fluo-
rescence signal is directly related to the distribution
of the molar concentration of the excited species, a
pixel-by-pixel computation involving species com-
binations is straightforward.
All possible two-species combinations (A1 , A2 )
associated to corresponding exponents (s1 , s2 ) have
been evaluated. The objective is to find a HRR
marker ([A1 ]s1 × [A2 ]s2 ) (where [Ai ] denotes the nor-
malized molar concentration of species number i) Fig. 1. Objective function q evaluated for OH × CH2 O
that minimizes the difference between the real heat with all possible exponents in Case F (for HHRR). Note
release profile and the reconstructed profile based that the z-axis is inverted, so that the best possible recon-
on the marker combination. struction (minimizing q) appears as a peak.
To quantify the difference between the heat re-
lease profile and the reconstructed profile, the di-
rect pixel-counting method described in [20] is used. 4. Results and discussion
First, the binarized image of normalized heat re-
lease [Hr ] obtained by DNS is thresholded between First, only the high heat release region
two user-chosen critical values crmin and crmax (with (0.5 ≤ [Hr ] ≤ 1) is considered. The objective func-
0 ≤ crmin < crmax ≤ 1). Then, all pixels with crmin tion q for all LIF-measurable species combinations
≤ [Hr ] ≤ crmax have a value of 1, all other pixels are using either unity exponents (s1 = s2 = 1) or opti-
set to 0. The binarized image for the reconstructed mal exponents have been computed for all the cases
profile is also treated in exactly the same way; the of Table 1. Due to length limitation, corresponding
pixels of the normalized field of ([A1 ]s1 × [A2 ]s2 ) results are shown in the supplemental materials.
with a value between crmin and crmax are set to 1, and Only selected conditions are discussed in what
the others to 0. Finally the number of pixels with a follows.
different value at the same location in both images First, best results from the literature, based on
are counted, leading to a value npix1 . The number HCO, (OH × CH2 O) and (H × CH2 O) are con-
of pixels with a value of 1 in the heat release images firmed by checking the corresponding objective
are counted as npix2 . The objective function – to be functions in Table S1 - S20; they show indeed
minimized – is finally defined as q = npix1 /npix2 , and a much lower value of q than all other possible
is a non-dimensional quantity describing the differ- species combinations involving unity exponents.
ence between both images. A value of q = 0 means Looking at Fig. 1, it can be seen that the re-
a perfect agreement. construction quality can be improved by introduc-
The user-defined critical thresholds crmin and ing additional exponents for a given species com-
crmax are used to decide which heat release region is bination. For instance, the objective function q
of particular interest for the comparison. In what for the conventional marker combination ([OH] ×
follows, the low heat release region and the high [CH2 O]), corresponding to unity exponents, is 5.9%
heat release region are considered separately. First, (lower arrow in Fig. 1), while it drops down to 3.0%
the high heat release region is investigated, choos- for optimal exponents, ([OH]1.09 × [CH2 O]1.18 ). The
ing crmin = 0.5 and crmax = 1.0. In most practical ap- superiority of markers involving optimal expo-
plications, the objective is to maximize heat release, nents is also obvious in Fig. 2, as a function of the
and regions with high heat release are obviously of equivalence ratio. There, it is shown that the recon-
particular interest. High heat release is also used to struction quality of conventional markers (unity
define the flame front [1]. On the other hand, the exponents, dashed curves) varies a lot as function
region of low heat release (defined here using crmin of the equivalence ratio. It is critical to obtain a
= 0.1 and crmax = 0.3) can be of particular interest generally valid marker of HRR with a low sensi-
when investigating flame quenching, and is consid- tivity towards equivalence ratio.
ered in a second step. For this purpose, a fitting function is introduced
Finally, the objective is to find the best combi- to obtain the best exponents (s1 , s2 ) for all flames at
nations of species measurable by LIF, and the as- different equivalence ratios ranging from 0.6 to 1.4,
sociated exponents, leading to a minimum value of by maximizing:
the objective function q for either high (HHRR) or
low heat release rate (LHRR). All possible expo- 1.4
1 − q(s1 , s2 )|φ=x
nent values s1 and s2 between 0 and 3 are systemat- f (s1 , s2 ) = , (3)
x=0.6
1 − qmin |φ=x
ically tested considering two digits, following [20].

Please cite this article as: C. Chi et al., DNS study of the optimal heat release rate
marker in premixed methane flames, Proceedings of the Combustion Institute (2018),
https://doi.org/10.1016/j.proci.2018.07.095
JID: PROCI
ARTICLE IN PRESS [m;August 4, 2018;5:21]

C. Chi et al. / Proceedings of the Combustion Institute 000 (2018) 1–9 5

The first one is basically simply [HCO], as orig-


inally identified in the seminal work of [8,9]. It is
known that it is very difficult to measure this quan-
tity, and it will not be discussed further.
Now, Cases J (with lower turbulence in-
tensity) and K (with higher turbulence inten-
sity) are performed to investigate the possi-
ble influence of varying turbulence intensity on
the identified markers. Exactly the same opti-
mal combinations are found in the same order
for both cases, demonstrating again the robust-
ness of these indicators. Checking now quanti-
tatively the performance of the proposed com-
binations, for Case J, q([H]1.11 × [CH2 O]0.86 ) =
6.5% and q([OH]1.07 × [CH2 O]1.17 ) = 2.8%; while
for Case K, q([H]1.11 × [CH2 O]0.86 ) = 7.3% and
q([OH]1.07 × [CH2 O]1.17 ) = 3.8%. These are excel-
lent results, documenting the quality of the recon-
struction of HHRR.
Fig. 2. Objective function q evaluated for conventional Going one step further concerning a possible
HRR markers with unity or with optimal exponents for impact of turbulence, laminar methane flames have
Cases A to I (for HHRR). Note that the z-axis is inverted,
so that the best possible reconstruction (minimizing q) ap-
been considered additionally for equivalence ratios
pears as a peak. The optimal exponents (denoted “opt”) of 0.7, 1.0 and 1.3 in Fig. 3. These laminar flames
are always those found for a specific equivalence ratio, and are simulated using GRI Mech 3.0 with the same
may thus vary from point to point. 2D spherical flame configuration, but without tur-
bulence. The established marker ([OH] × [CH2 O])
(unity exponents, dashed lines) is highly sensitive
where qmin |φ=x is the lowest objective function ob- to the equivalence ratio, much more than the al-
tained at a given equivalence ratio φ = x. This pro- ternative combination [H] × [CH2 O], confirming
cedure also covers the case of single species (like the findings of [17]. Introducing now our optimal
HCO), when s2 = 0. exponent combinations, it is observed that the re-
The HCO-based markers (including [HCO], constructions are always better, increasingly so for
([HCO] × [OH]), ([HCO] × [CH2 O]) and increasing equivalence ratio, and are much less sen-
([HCO] × [H])) are first checked. Unfortunately, it sitive to the equivalence ratio (solid lines in Fig. 3).
is found for the last three combinations that the Overall, [OH]1.07 × [CH2 O]1.17 is slightly superior
best exponents for the second member of the com- for these laminar flames.
bination (either [OH], [CH2 O], or [H]) are close The difficulty of finding an accurate and
to 0 and vary in a stochastic manner for different generally-valid HRR marker is again obvious when
equivalence ratios. This shows that [HCO] alone looking at Fig 3; the additional exponents are help-
is indeed a better marker than in any combina- ful for this purpose. For a laminar flame with
tion. The next candidates are [HCO] (kept in the φ = 1.0, the reconstruction [H] × [CH2 O] (dashed
analysis), ([H] × [CH2 O]) and ([OH] × [CH2 O]). line without symbols in Fig 3(b) shows a relatively
Finally, considering the GRI Mech 3.0 mechanism, poor reconstruction quality; the objective function
the best HHRR markers at all equivalent ratios is here q = 16%. The reconstruction does not pre-
are found to be [HCO]1.09 , closely followed by dict correctly the location of the peak HRR region,
([H]1.13 × [CH2 O]0.84 ) and ([OH]1.2 × [CH2 O]1.24 ). with a deviation exceeding 10%. The proportional-
For the San Diego mechanism, the best mark- ity is also poor, the dashed line being quite far from
ers are [HCO]0.96 , ([H]1.09 × [CH2 O]0.89 ) and the diagonal. However, the same reconstruction
([OH]0.93 × [CH2 O]1.09 ). It can be seen that the with optimal exponents [H]1.11 × [CH2 O]0.86 shows
optimal exponents for all optimal combinations a much better correlation, with q = 7.5% (solid line
are not identical for both reaction mechanisms, without symbols in Fig 3(b). The corresponding
but quite similar. Involving now in the analysis all prediction lies very close the diagonal, and is thus a
results obtained for all different equivalence ratios good approximation. Comparing these two cases, a
and both reaction mechanisms (Cases A - I and reconstruction with an objective function q < 10%
A0 - I0), the best solutions in general are finally is considered in this study as the necessary thresh-
found to be: old for high-quality reconstruction, with even lower
values of q indicating higher reconstruction accu-
[HCO]1.01 , racy.
([H]1.11 × [CH2 O]0.86 ), Turning back to turbulent flames, the joint
probability density functions (PDF) of the corre-
([OH]1.07 × [CH2 O]1.17 ). lations between the normalized HRR and the re-

Please cite this article as: C. Chi et al., DNS study of the optimal heat release rate
marker in premixed methane flames, Proceedings of the Combustion Institute (2018),
https://doi.org/10.1016/j.proci.2018.07.095
JID: PROCI
ARTICLE IN PRESS [m;August 4, 2018;5:21]

6 C. Chi et al. / Proceedings of the Combustion Institute 000 (2018) 1–9

Fig. 3. Normalized HRR vs. reconstruction from (a) [OH] × [CH2 O] and (b) [H] × [CH2 O], with unity or with optimal
exponents for laminar flames at different equivalence ratios.

Fig. 4. Joint PDF of the correlation between normalized HRR and (a) reconstruction from ([OH] × [CH2 O]);
(b) reconstruction from ([OH]1.07 × [CH2 O]1.17 ); (c) reconstruction from ([H] × [CH2 O]) or (d) reconstruction from
([H]1.11 × [CH2 O]0.86 ).

Please cite this article as: C. Chi et al., DNS study of the optimal heat release rate
marker in premixed methane flames, Proceedings of the Combustion Institute (2018),
https://doi.org/10.1016/j.proci.2018.07.095
JID: PROCI
ARTICLE IN PRESS [m;August 4, 2018;5:21]

C. Chi et al. / Proceedings of the Combustion Institute 000 (2018) 1–9 7

Fig. 5. Case M (3D DNS): Clipped iso-surface of (a) normalized heat release; (b) conventional reconstruction
by ([OH] × [CH2 O]) with unity exponents; (c) reconstruction by ([OH]1.07 × [CH2 O]1.17 ); (d) reconstruction by
([H]1.11 × [CH2 O]0.86 ). The iso-surfaces are clipped with 3 cut-planes normal to x, y and z directions, respectively.

constructions are shown in Fig. 4. The joint PDF is exponents proposed previously (see Fig. 5(c) and
calculated using all the simulation results obtained (d)) reproduce perfectly the high heat release region
with GRI Mech 3.0, for an equivalence ratio φ = (Fig. 5(a). Using unity exponents (see Fig. 5(b), no-
0.6−1.4 (Cases A to I). This figure demonstrates ticeable discrepancies appear in the high heat re-
again the positive impact of the exponents intro- lease region.
duced together with the species combinations. The Finally, it is possible to apply exactly the same
joint PDF in Fig. 4(b) and (d) lie much closer to the analysis to the low heat release region (LHRR).
diagonal compared with the conventional markers This would be particularly useful for analyzing ex-
involving unity exponents (shown in Figs. 4(a) and tinction processes, or quenching followed by re-
(c)). ignition. Using a similar analysis, the optimized
All the previous studies relied on 2D DNS. It marker combinations for the low heat release re-
is still impossible to carry out such a systematic gion (0.1 ≤ [Hr ] ≤ 0.3) are found to be:
study in 3D with the available computing power.
Nevertheless, a single 3D case has been computed
with the reduced methane mechanism of Smooke [HCO]0.89 ,
[26] (Case M), in order to check possible 3D effects.
([H]1.14 × [CH2 O]0.86 ),
As shown by Fig. 5, the reconstruction of heat re-
lease using the optimal marker combinations and ([OH]0.95 × [CH2 O]1.0 ).

Please cite this article as: C. Chi et al., DNS study of the optimal heat release rate
marker in premixed methane flames, Proceedings of the Combustion Institute (2018),
https://doi.org/10.1016/j.proci.2018.07.095
JID: PROCI
ARTICLE IN PRESS [m;August 4, 2018;5:21]

8 C. Chi et al. / Proceedings of the Combustion Institute 000 (2018) 1–9

It is interesting to see that these three combina- References


tions, and their order of appearance, are identical
to that found for HHRR. Even the exponents do [1] C. Zistl, R. Hilbert, G. Janiga, D. Thévenin, Comput.
not change much. Overall, the indicators derived to Vis. Sci. 12 (2009) 383–395.
describe the high heat release region can be used to [2] W. Hubschmid, R. Bombach, A. Inauen, et al., Exp.
Fluids 45 (1) (2008) 167–182.
investigate as well extinction phenomena.
[3] Y. Hardalupas, M. Orain, Combust. Flame 139 (3)
(2004) 188–207.
[4] K. Zähringer, D. Durox, F. Lacas, Int. J. Heat Mass
Transf. 46 (2003) 3539–3548.
5. Conclusions and perspectives
[5] C. Panoutsos, Y. Hardalupas, A. Taylor, Combust.
Flame 156 (2009) 273–291.
In the present study, premixed spherical [6] A. Gazi, G. Vourliotakis, G. Skevis, M.A. Founti,
methane-air flames have been simulated by Direct Combust. Sci. Tech. 185 (2013) 1482–1508.
Numerical Simulations. A pixel-to-pixel image [7] N.B. Anikin, R. Suntz, H. Bockhorn, Appl. Phys. B
comparison is used to quantify the reconstruction 107 (2012) 591–602.
quality of different HRR markers and obtain [8] H.N. Najm, P.H. Paul, C.J. Mueller, P.S. Wyckoff,
optimal combinations, taking also into account an Combust. Flame 113 (1998) 312–332.
exponent for each species concentration. Optimal [9] P.H. Paul, H.N. Najm, Proc. Combust. Inst. 27 (1998)
43–50.
HRR markers have been found, able to describe
[10] A. Fayoux, K. Zähringer, O. Gicquel, J.C. Rolon,
premixed flames at an equivalence ratio ranging Proc. Combust. Inst. 30 (2005) 251–257.
from 0.6 to 1.4, either laminar or turbulent, both [11] R.L. Gordon, A.R. Masri, E. Mastorakos, Combust.
in 2D and 3D configurations, using two different Theory Model. 13 (4) (2009) 645–670.
reaction mechanisms, and describing both high [12] S. Pfadler, F. Beyrau, A. Leipertz, Opt. Express 15
and low heat release levels. (23) (2007) 15444–15456.
Apart from [HCO] used as single species, [13] M. Röder, T. Dreier, C. Schulz, Appl. Phys. B 107 (3)
but barely measurable, the optimal marker (2012) 611–617.
combinations measurable by LIF are [14] M. Röder, T. Dreier, C. Schulz, Proc. Combust. Inst.
34 (2013) 3549–3556.
([OH]1.07 × [CH2 O]1.17 ) and ([H]1.11 × [CH2 O]0.86 ).
[15] S. Böckle, J. Kazenwadel, T. Kunzelmann, D.-I. Shin,
Those are found to be noticeably better than the C. Schulz, J. Wolfrum, Proc. Combust. Inst. 28 (2000)
same species combinations with unity exponents. 279–286.
The combination ([H]1.11 × [CH2 O]0.86 ) is a [16] Z.M. Nikolaou, N. Swaminathan, Combust. Flame
slightly better choice to predict heat release over 161 (12) (2014) 3073–3084.
its whole range, even for quenching flames. How- [17] I.A. Mulla, A. Dowlut, T. Hussain, et al., Combust.
ever, [H] can currently only be measured with the Flame 165 (2016) 373–383.
two-photon excitation process. As an alternative, [18] Z. Wei, C. Leung, C. Cheung, Z. Huang, Energy 133
([OH]1.07 × [CH2 O]1.17 ] is a very good compromise, (2017) 35–45.
[19] W.D. Kulatilaka, S. Roy, N. Jiang, J.R. Gord, Appl.
allowing relatively easy measurements while still
Phys. B 122 (2) (2016) 26.
leading to a very good prediction. [20] C. Chi, G. Janiga, A. Abdelsamie, K. Zähringer,
T. Turányi, D. Thévenin, Flow Turbul. Combust. 98
(4) (2017a) 1117–1132.
[21] C. Chi, G. Janiga, K. Zähringer, D. Thévenin, VDI
Acknowledgments Ber. 2302 (2017b) 625–636.
[22] A. Abdelsamie, G. Fru, T. Oster, F. Dietzsch,
The financial support of the International Max G. Janiga, D. Thévenin, Comput. Fluids 131 (2016)
Planck Research School Magdeburg for Advanced 123–141.
Methods in Process and Systems Engineering (IM- [23] A. Abdelsamie, D. Thévenin, Proc. Combust. Inst. 36
PRS ProEng) for C. Chi is gratefully acknowl- (2) (2017) 2493–2502.
edged. The computer resources provided by the [24] G. Smith, D. Golden, M. Frenklach, et al., available
Gauss Center for Supercomputing/Leibniz Super- at http://www.me.berkeley.edu/gri_mech/, 2017.
computing Center Munich have been absolutely [25] Chemical-Kinetic Mechanisms for Combustion Ap-
plications, San Diego Mechanism web page, Me-
necessary to carry out this project. Scientific discus- chanical and Aerospace Engineering (Combustion
sions with A. Abdelsamie are gratefully acknowl- Research), University of California at San Diego,
edged. available at: http://combustion.ucsd.edu/, 2017.
[26] M. Smooke, V. Giovangigli, M. Smooke, in: Lecture
Notes in Physics, Berlin Springer Verlag, 384, 1991.
[27] D.G. Goodwin, H.K. Moffat, R.L. Speth, available
Supplementary material at: http://www.cantera.org/, 2015.
[28] D. Thévenin, F. Behrendt, U. Maas, B. Przywara,
Supplementary material associated with this ar- J. Warnatz, Comput. Fluids 25 (5) (1996) 485–496.
ticle can be found, in the online version, at doi:10. [29] N. Peters, Turbulent Combustion, Cambridge Univer-
1016/j.proci.2018.07.095. sity Press, Cambridge, 2000.

Please cite this article as: C. Chi et al., DNS study of the optimal heat release rate
marker in premixed methane flames, Proceedings of the Combustion Institute (2018),
https://doi.org/10.1016/j.proci.2018.07.095
JID: PROCI
ARTICLE IN PRESS [m;August 4, 2018;5:21]

C. Chi et al. / Proceedings of the Combustion Institute 000 (2018) 1–9 9

[30] C.J. Rutland, A. Trouvé, Combust. Flame 94 (1–2) [34] W.D. Kulatilaka, J.H. Frank, T.B. Settersten, Proc.
(1993) 41–57. Combust. Inst. 32 (1) (2009) 955–962.
[31] B.E. Battles, R.K. Hanson, J. Quant. Spectrosc. Ra- [35] B. Ayoola, R. Balachandran, J. Frank, E. Mas-
diat. Transfer 54 (3) (1995) 521–537. torakos, C. Kaminski, Combust. Flame 144 (1) (2006)
[32] P.H. Paul, J.A. Gray, J.L. Durant, J.W. Thoman, 1–16.
Appl. Phys. B 57 (4) (1993) 249–259.
[33] N. Garland, D. Crosley, Proc. Combust. Inst. (1986)
1693–1702.

Please cite this article as: C. Chi et al., DNS study of the optimal heat release rate
marker in premixed methane flames, Proceedings of the Combustion Institute (2018),
https://doi.org/10.1016/j.proci.2018.07.095

You might also like