You are on page 1of 11

Fuel 259 (2020) 116152

Contents lists available at ScienceDirect

Fuel
journal homepage: www.elsevier.com/locate/fuel

Full Length Article

Laminar burning velocities of CH4/O2/N2 and oxygen-enriched CH4/O2/CO2 T


flames at elevated pressures measured using the heat flux method

Shixing Wanga, Zhihua Wanga, , Yong Hea, Xinlu Hana, Zhiwei Sunb, Yanqun Zhua, Mario Costac
a
State Key Laboratory of Clean Energy Utilization, Zhejiang University, Hangzhou 310027, China
b
School of Mechanical Engineering and Centre for Energy Technology, The University of Adelaide, Australia
c
IDMEC, Mechanical Engineering Department, Instituto Superior Técnico, Universidade de Lisboa, Lisboa, Portugal

A R T I C LE I N FO A B S T R A C T

Keywords: Laminar burning velocities (SL) of CH4/O2/N2 and oxygen-enriched CH4/O2/CO2 flames were measured at
Laminar burning velocity elevated pressures up to 0.5 MPa and equivalence ratios ranging from 0.6 to 1.6. The oxygen molar fraction was
Heat flux method varied from 0.18 to 0.23 in the O2/N2 mixtures and from 0.31 to 0.42 in the O2/CO2 mixtures. The experimental
High pressure results showed good agreement with the results reported in previous works, validating the suitability and re-
Methane
liability of the present experimental method for measuring SL at high pressure. Kinetic modelling was also
Oxygen concentration
performed using the GRI-Mech 3.0 and the HP-Mech mechanisms. Both mechanisms predict reasonably well SL
N2- and CO2-dilution
and a power factor β that quantifies the dependence of SL on pressure. Thermal-diffusion effects play a major role
in the laminar burning velocity decrease due to CO2 dilution at normal and elevated pressures. Kinetic analysis
indicated that the reverse of reaction CO + OH = CO2 + H retards the flame propagation in competition with
H + O2 = O + OH. Competition of the H consuming reaction H + O2 = O + OH with the two CH3 consuming
reactions 2CH3 (+M) = C2H6 and CH3 + H (+M) = CH4 (+M) leads to a non-monotonic behavior of the
overall reaction order for both the N2- and CO2-diluted flames.

1. Introduction needed for validating and developing reliable combustion mechanisms


for OEC with N2, CO2 and water vapor dilution.
Oxygen-enriched combustion (OEC) is a promising approach for Table 1 summarizes previous studies on the effect of N2 and CO2
highly efficient power generation not only because of the low NOx dilution on the laminar burning velocities of oxygen-enriched methane
emissions but also because of the suitability for CO2 capture and storage flames performed at atmospheric and elevated pressures. The studies
[1,2]. In OEC processes, fuel can be burned in air or in a mixture of undertaken at atmospheric pressure consistently revealed that dilution
recycled flue gas with pure oxygen. Therefore, gases like N2, CO2 and with both CO2 and N2 reduce the burning velocities, with the CO2
water vapor are usually involved as dominant components in the oxi- producing a more significant impact owing to thermal and chemical
dation flows [3]. To efficiently utilize OEC and to develop reliable effects. These conclusions still need validation at elevated pressures
predictive models, it is necessary to understand the combustion char- owing to the very limited studies under these conditions. De Persis et al.
acteristics of fuels at elevated pressures in these diluted gases. Among [6] investigated experimentally and numerically the laminar burning
the parameters required for knowledge enhancement, a key parameter velocities of CO2-diluted CH4/O2/N2 flames with oxygen concentration
is the laminar burning velocity (SL), which is closely relevant to flame varied from 0.21 to 1.0 in the oxidizer stream (O2 + N2). They found
stability, diffusion probability and heat release rate. However, most of that the addition of CO2 reduces SL significantly for pressures between
the previous investigations were conducted at atmospheric pressure. 0.1 and 0.8 MPa. Xie et al. [7] investigated the effect of CO2 dilution on
This represents an important research gap for many practical OEC SL at pressures up to 0.3 MPa and also concluded that CO2 inhibits SL
processes since most of them are operated at elevated pressures, e.g., through the consumption of the H radical by the reverse of reaction
pressurized boilers and gas turbines [4,5]. Indeed, experimental result CO + OH = CO2 + H in competition with the reaction
of SL for OEC at elevated pressures is rather scarce, even for methane, H + O2 = O + OH, as reported by Glarborg et al. [8].
which is the dominant component of natural gas. Therefore, compre- Table 1 also reveals that SL measurements at elevated pressure were
hensive experimental flame speed data at elevated pressures are much often performed using the outwardly propagating flame method. In this


Corresponding author.
E-mail address: wangzh@zju.edu.cn (Z. Wang).

https://doi.org/10.1016/j.fuel.2019.116152
Received 26 July 2019; Received in revised form 2 September 2019; Accepted 4 September 2019
0016-2361/ © 2019 Elsevier Ltd. All rights reserved.
S. Wang, et al. Fuel 259 (2020) 116152

Table 1
Previous studies on the effects of N2 and CO2 dilution on laminar burning velocities of oxygen-enriched methane flames.
Reference Method* Dilution gases O2 concentration P (MPa) ϕ

Hu et al. [17] CF CO2 25–35% 0.1 0.6–1.4


Mazas et al. [18] CF CO2 20–40% 0.1 0.4–1.6
Jeongseog et al. [19] CF 100% 0.1 0.5–2.0
Dirrenberger et al. [11] HFM N2 21–35% 0.1 0.4–2.1
Konnov et al. [20] HFM CO2 26–35% 0.1 0.6–1.6
Chen et al. [21] OPF CO2 30% 0.1–0.5 0.5–0.8
Khan et al. [22] OPF CO2, N2 18–54% 0.1 0.6–1.4
De Persis et al. [6] OPF CO2, N2 20–100% 0.1–0.8 0.7–1.1
Xie et al. [7] OPF CO2 30–60% 0.1–0.3 0.4–1.6
Cai et al. [23] OPF N2 21–60% 0.1–0.5 0.5–1.5

* OPF – outwardly propagating flame, CF – conical flame, HFM – heat flux method.

context, flame speed data for OEC at elevated pressures measured with
other methods are needed for cross-validation, particularly since it is
known that the spherical flame method is affected by the stretch effect
at elevated pressures. The heat flux method has been widely used at
atmospheric pressure [9–13], but only by two groups [14–16] at ele-
vated pressures. Therefore, it is very useful and beneficial to expand the
application of the heat flux method to OEC at elevated pressures owing
to the important advantages of the method, such as the negligible in-
fluences of heat loss and flame stretch.
In light of the discussion above, the specific aims of the present
work are (1) to expand the application of the heat flux method at ele-
vated pressures and thereby to gather highly reliable SL data for CH4/
O2/N2 flames up to 0.5 MPa and oxygen-enriched CH4/O2/CO2 flames
up to 0.3 MPa; (2) to quantify the thermal-diffusion and chemical ef-
fects of CO2 dilution on SL; and (3) to examine the pressure dependence
of SL on the equivalence ratio.

2. Experiments and simulations

2.1. Heat flux method at high pressure

The heat flux method introduced by Bosschaart and de Goey [24,25]


has been widely applied for SL measurements at atmospheric pressure
[9–13]. Details of the heat flux method and associated uncertainties can
be found elsewhere [9,11,26–28]. In the heat flux method, the heat loss
from the flame to the burner can be compensated by the heat gain of the
unburnt gas flowing through the burner plate. When the radial tem- Fig. 1. Schematic of the high-pressure chamber fitted with a heat flux burner.
perature gradient is equal to zero, the adiabatic laminar burning velo-
city can be determined from the gas feeding velocity. Briefly, the ad- the chamber. It was found that the pressure difference between the top
vantage of the heat flux method is to establish a 1-D, adiabatic and and the bottom of the chamber could be 0.0025 MPa. Therefore, a
stretch-less flat flame. To the best of our knowledge, only few works pressure sensor was located at the height of the burner surface to make
applied this method at elevated pressures [14–16]. accurate inner pressure measurements.
Fig. 1 shows a schematic of the high-pressure chamber fitted with a Fig. 2 shows the entire high-pressure facility for measuring laminar
heat flux burner. The heat flux burner has been modified and is dif- burning velocities. The gases were supplied to the burner through mass
ferent from the one we used previously at atmospheric pressure [9]. flow controllers (Brooks Instruments Co., Ltd., 5850E). The pressure
The diameter of the perforated plate is 20 mm, instead of 30 mm, to was controlled by a parallel combination of a manually operated valve
better handle the increased gas flow rates at elevated pressures. The and an electrical PID controller, which were installed in the exhaust gas
diameter and pitch of the holes on the perforated plate were reduced pipe. The exhaust gases were released from the top of the chamber and
from 0.5 to 0.3 mm and from 0.7 to 0.4 mm, respectively. These re- forced through a heat exchanger and a condenser to remove the
ductions help to minimize local flame stretch and to avoid the cellular moisture from the gases before the pressure valves. Gases with high
instability of the flames at elevated pressures [29], Moreover, the entire purities (CO2, 99.99%; N2, 99.99%; O2, 99.99%; CH4, 99.95%) were
plenum chamber of the burner was filled with small stainless-steel used in this study.
beads (3 mm in diameter) to enhance the mixing of the gases supplied.
The high-pressure chamber has a volume of 35 L with a maximum
sustained pressure of 5 MPa. Such a large volume is beneficial in 2.2. Assessment of the measurement uncertainties
keeping the inner pressure constant. Four optical windows were de-
signed and installed for flame observation and optical measurements. A Alekseev et al. [30] assessed the experimental uncertainties asso-
nitrogen flow was used to purge each optical window to avoid moisture ciated with SL measurements obtained with the heat flux method at
condensation and particle deposition. Above the burner surface, mul- atmospheric pressure. A similar analysis was performed in the present
tiple cooling coils were installed to stabilize the temperature inside the work for elevated pressures. The two dominant uncertainty sources are
chamber. Pressure and temperature sensors were also installed inside those associated with the temperatures of the burner plate measured

2
S. Wang, et al. Fuel 259 (2020) 116152

Fig. 2. High pressure facility for laminar burning velocity measurement by heat flux method.

using multiple thermocouples and those associated with the inlet ve- uncertainty due to the pressure controllers used in the present setup,
locities of gases, which are related with the uncertainties of the mass which can be calculated as follows:
flow controllers.
ΔP
SL is determined by interpolating the parabolic coefficient C that ΔSLP = −SL (1 + β')
P set (6)
approximates the temperature distribution of the burner plate to 0 at
different inlet velocities, u. The measurement sensitivity s is determined where ΔP is the uncertainty of the pressure controllers and β′ is fitted by
as follow: the commonly accepted power law:
dC '
s= |u = SL
(1) SLm/ SLset = (P m/ P set ) β (7)
du
The uncertainty of SL associated with the temperature measure- where SLm, set m set
SL , p , p are the actual and the set values of SL and
ments of the thermocouples can be calculated as follow: pressure, respectively. The overall uncertainty of the measured SL can
therefore be evaluated based on the three uncertainties discussed
σC
ΔSLTC = above.
s (2)
Fig. 3 presents the typical values of ΔSL for stoichiometric methane/
where σc is the average standard deviation of the thermocouple tem- air flames at different pressures and equivalence ratios. Fig. 3a shows
peratures, which has a specific value for a given burner [28]. σc can be that, at elevated pressures, the measurement uncertainty decreases with
calculated as follows [30]:
1/2
⎛ ∑ (Ti − C∙ri 2 − Tcenter )2 ⎞
σc = ⎜ − ⎟
2 2 2
⎝ (N − 2) ∙ ∑ (ri − r ) ⎠ (3)

where N is the number of thermocouples, and Ti are the measured


temperatures at radii ri (i = 1, 2, …, 8 as eight thermocouples were
used).
The uncertainties associated with the mass flow controllers can be
estimated as follows:

( ∑ (ΔFi )2)1/2
ΔSLMFC = SL
Ftot (4)

where Ftot is the total flow rate of all gases, and ΔFi is the uncertainty of
the ith mass flow controller. ΔFi can be calculated as follows:
ΔFi = 0.8%∙Rd + 0.2%∙Fs (5)

where Rd is the measured mass flow and Fs is the full range of the mass
flow controller. Fig. 3. Typical values of measurement uncertainty ΔSL for the CH4/air flames
For SL measurements at elevated pressures, there is an additional (a) at ϕ = 1.0 and (b) P = 0.3 MPa.

3
S. Wang, et al. Fuel 259 (2020) 116152

the increase of the pressure; specifically, it is 0.64 cm/s at 0.3 MPa and
0.39 cm/s at 0.5 MPa. Fig. 3b shows that, at 0.3 MPa, the uncertainty of
SL is little sensitive to the equivalence ratio, with a typical value of
~0.5 cm/s.

2.3. Kinetic simulations

Kinetic simulations of laminar burning velocity were performed for


the flames studied experimentally. The software CHEMKIN PRO was
used and the one-dimensional-freely-propagating laminar flame speed
model (PREMIX) was adopted in simulations. Two kinetic mechanisms
were used: the GRI-Mech 3.0 [31] mechanism and HP-Mech [32] me-
chanism.
An adaptive re-gridding method with convergence conditions of
GRAD and CURV values setting at 0.02, which ensures at least 500 grid
points, was used for each test condition. Thermal diffusion (Soret effect)
and mixture-averaged transport option were used in the computations.

3. Results and discussion

3.1. Effects of equivalence ratio and pressure on SL of CH4/O2/N2 flames

Fig. 4 shows the measured and predicted laminar burning velocities


of CH4/O2/N2 flames as a function of equivalence ratio for three oxygen
molar fractions at 0.1, 0.3 and 0.5 MPa. It is clear that a small increase
in the oxygen content significantly enhances the burning velocities for
both atmospheric and elevated pressures. As the pressure increase, i.e.,
the gas density, SL decreases significantly. It should be pointed out that
values of SL obtained in this work agree well with those available in the
literature [13,14,27,33–38]. Fig. 4 also shows that the two kinetic
mechanisms used in this study (GRI-Mech 3.0 and HP-Mech) reproduce
well the present experimental data. All these observations corroborate
the suitability of the heat flux method to measure flame speeds at
elevated pressures. It should be noted that the oxygen molar fraction of
these CH4/O2/N2 flames varied slightly around 0.21, which do not
correspond to OEC conditions, but these flames have similar SL values
and adiabatic flame temperatures of the CH4/O2/CO2 flames studied
(see sections below), providing a good validation of the present ex-
perimental set-up, before examining the CO2-diluted flames.
Fig. 5 shows the laminar burning velocities of CH4/O2/N2 flames as
a function of the pressure for three oxygen molar fractions at ϕ = 1.0,
and the respective fitting curve based on the empirical power law, i.e.,
SL/SL0 = (p/p0)β, where β represents the overall reaction order [39]. It
is found that the empirical power law fits well the experimental data,
originating β values of −0.482, −0.393 and −0.373 for oxygen molar
fractions of 0.18, 0.21 and 0.23, respectively. The values of β evaluated
from the predicted SL data are slightly inferior; specifically, −0.499,
−0.428 and −0.394 for the three oxygen contents, respectively, based
on the results predicted by the GRI-Mech 3.0, and −0.483, −0.397 and Fig. 4. Measured and predicted laminar burning velocities of CH4/O2/N2
flames as a function of the equivalence ratio for three oxygen molar fractions at
−0.353 based on the results predicted using the HP-Mech. Although
0.1, 0.3 and 0.5 MPa. Solid symbols: present study, open symbols: literature
quasi-linear relationships are observed in Fig. 5, it should be noted that
data [13,14,27,33–38], solid lines: GRI-Mech 3.0, dash lines: HP-Mech.
they do not exactly follow the power law due to the slight decrease of
the overall reaction order as pressure increases (see Fig. 6), which is in
line with previous studies [33,39]. Nonetheless, the overall reaction 3.2. Effects of equivalence ratio and pressure on SL of CH4/O2/CO2 flames
order can be assumed constant in the pressure range 0.1–0.5 MPa to
facilite the analysis as suggested in references [14,40]. Fig. 8 shows the measured and predicted laminar burning velocities
Fig. 7 shows the values of β, obtained from measurements and of CH4/O2/CO2 flames as a function of the equivalence ratio for three
predictions, as a function of the equivalence ratio for the three oxygen oxygen molar fractions at 0.1 and 0.3 MPa. The dependences of SL on
contents. The β values of 21% O2 condition fits well with data of Gos- oxygen concentration and pressure are consistent with those observed
wami [14] and Gu [36], while data of Halter [41] is lower than the earlier for the N2-diluted flames. The CH4/O2/CO2 flames, however,
present study. It is seen that the value of β reaches a maximum around they present larger discrepencies between the measured and predicted
the stoichiometric condition, followed by a complex dependence for the SL data. It is verified that the GRI-Mech 3.0 predicts better the experi-
fuel-rich conditions. This complex dependence of β on equivalence ratio mental data than the HP-Mech for fuel-lean conditions, but the latter
is discussed below. As for the dependence of β on oxygen content, the mechanism does better for fuel-richconditions.
former decreases as the latter decreases. Fig. 9 shows laminar burning velocities of CH4/O2/CO2 flames as a

4
S. Wang, et al. Fuel 259 (2020) 116152

Fig. 5. Laminar burning velocities of CH4/O2/N2 flames as a function of the


pressure for three oxygen molar fractions at ϕ = 1.0, and the respective fitting
curves. Symbols: present study, lines: fitting curve. The factor β represents the
overall reaction order [39].

Fig. 8. Measured and predicted laminar burning velocities of CH4/O2/CO2


flames as a function of the equivalence ratio for three oxygen molar fractions at
0.1 and 0.3 MPa. Solid symbols: present study, open symbols: literature data
[17]. Solid lines: GRI-mech, dash lines: HP-mech.
Fig. 6. Calculated overall reaction order, n, for CH4/O2/N2 and CH4/O2/CO2
flames as a function of the pressure for three oxygen molar fractions at ϕ = 1.0.

Fig. 9. Laminar burning velocities of CH4/O2/CO2 flames as a function of


pressure for three oxygen molar fractions at ϕ = 1.0, and the respective fitting
Fig. 7. Measured and predicted values of β for CH4/O2/N2 flames, obtained
curves. Symbols: present study, lines: fitting curve. The factor β represents the
form measurements and predictions, as a function of the equivalence ratio for
overall reaction order [38].
the oxygen contents. Solid symbols: present study, open symbols: literature data
[14,36,41], lines: GRI-Mech 3.0.
but the values of β are now larger than those for the N2- diluted flames
function of the pressure for three oxygen molar fractions at ϕ = 1.0 and with lower oxygen content (cf. Fig. 7). For the CO2- diluted flames, the
the respective fitting curve, while Fig. 10 shows the values of β, ob- values of β are −0.357, −0.28 and −0.237 for oxygen molar fractions
tained from measurements and predictions, as a function of the of 0.31, 0.38 and 0.42, respectively. As for the values of β evaluated
equivalence ratio for the three oxygen contents. It is seen that the trends from the predicted SL data, the GRI-Mech 3.0 originated −0.398,
are like those encountered for the N2- diluted flames (cf. Figs. 5 and 7), −0.307 and −0.273, and the HP-Mech −0.399, −0.302 and −0.269
for the three oxygen contents, respectively.

5
S. Wang, et al. Fuel 259 (2020) 116152

Fig. 10. Measured and predicted values of β for CH4/O2/CO2 flames, obtained
from measurements and predictions, as a function of the equivalence ratio for
the three oxygen contents, Symbols: present study, lines: GRI-Mech 3.0.

The presence of CO2 in the flame affects the accuracy of the SL


measurements, especially for elevated pressures due to the increased
radiative heat transfer. Chen et al. [21,42] concluded that the optical
thin model (OTM) over-predicted the radiative heat losses and thus
underpredicted the SL values of CO2 diluted CH4-O2 flames, and de-
veloped a fitted statistical narrow band (SNB) model to quantify the
radiation absorption effect in outwardly propagating spherical flames.
They found that radiation-induced uncertainties in the SL measure-
ments are nearly neglibible (within 2.5%) for CO2-diluted CH4-O2
flames at both normal and elevated pressures [42].
As for heat flux method, Naucler et al. [43] modeled the radiative
heat losses of CH3OH/O2/CO2 flames using the OTM and obtained a
reduction in SL of about 0.5 cm/s without considering the burner plate
reflection and re-absorption of the fresh gas mixture. Bosschaart et al.
[44] used CARS thermometry in the post-flame zone of a heat flux
burner and found that radiative losses affect the temperature profile,
but have no significant influence on the SL values of CH4/air flames Fig. 11. Contribution of the thermal-diffusion and chemical effects of the CO2
[27]. Moreover, as the pressure increases, the distance between the dilution at 0.1 and 0.5 MPa.
flame and the burner plate decreases, or, in other words, the peak heat
release zone of the 1-D flame shifts to upstream, which means that the
contribution was calculated as: Su [FCO2] − Su [CO2 ] , being their sum equal
heat exchange from the flame to the unburnt gas mixture is more fa- Su [FN2] − Su [CO2 ]
to 1.0.
vored than the heat losses from the flame to the surroundings. Conse-
Fig. 11(b, c) shows that the reduction of SL is mostly due to the
quently, the radiation-induced uncertainty in the SL measurements can
thermal-diffusion effect, more than 50% for CO2 dilution up to 60% in
be neglected in the present study.
the oxidizer. The chemical contribution increases as the O2 content
increases and reaches its maximum at ϕ = 1.1, which corresponds to
the peak SL location. The chemical effect decreases as the flame shifts to
3.3. Contribution of thermal-diffusion and chemical effects on SL of CH4/
the lean and rich sides, and increases again at the over-rich side. As the
O2/CO2 flames
pressure increases, the thermal-diffusion effect is enhanced, while the
chemical effect behaves oppositely due to the decrease of the overall
As previously investigated, CO2 and N2 dilution have three different
reaction rate as the pressure increases, counteracting the increased
effects on flame speed, namely the dilution effect, the thermal-diffusion
third-body collision effect of the CO2.
effect and the chemical effect [45–47]. The N2 dilution does not affect
The adiabatic flame temperature is a key parameter to represent the
the heat capacity and the thermal diffusivity of the mixture, and
heat release capacity and is strongly related to SL. Fig. 12 shows the
practically does not affect the laminar burning velocity through che-
laminar burning velocity as a function of the adiabatic flame tem-
mical reactions. Therefore, the same content of N2 and FCO2 (fictitious
perature, Tad, for pressures between 0.1 and 0.5 MPa. In addition to
CO2) were used to distinguish the thermal-diffusion and chemical ef-
three cases of different O2/(O2 + CO2) ratios (31%, 38% and 42%),
fects. FCO2 has exactly the same thermal and transport characteristics
three cases of 21% O2/(O2 + N2), 31% O2/(O2 + FCO2) and 31% O2/
of the CO2, but does not participate in any reaction [48]. Fig. 11 shows
(O2 + N2), which have the same Tad, the same thermal diffusivity and
the contribution of the thermal-diffusion and chemical effects of the
the same O2 content as 31% O2/(O2 + CO2), respectively, are also
CO2 dilution at 0.1 and 0.5 MPa. The difference between the N2 dilution
presented. As the pressure increases, SL decreases, showing a very weak
and the FCO2 dilution lines in Fig. 11(a) is caused by the difference
and even inverse correlation with Tad. The adiabatic flame temperature
between of the N2 and CO2 thermal-diffusion properties; and the dif-
decreases less markedly for the N2-diluted flames than for the CO2-di-
ference between the FCO2 dilution and CO2 dilution lines is caused by
luted flames. Obviously, it is not the flame temperature that leads to the
the chemical effect of the CO2 (the dilution effect represented by the
decrease of the SL as the pressure increases. The 31% O2/(O2 + CO2)
line of CH4-pure O2 and CH4-O2-N2 is not shown). The thermal-diffu-
and 21% O2/(O2 + N2) lines have the same Tad, 2200–2250 K, but
sion contribution was calculated as: Su [FN2] − Su [FCO2 ] and the chemical
Su [FN2] − Su [CO2 ]

6
S. Wang, et al. Fuel 259 (2020) 116152

R38 H+O2=O+OH
(a)

R119 HO2+CH3=OH+CH3O

R99 OH+CO=H+CO2 CH4-O2-CO2


= 1.0
R166 HCO+H2O=H+CO+H2O

153 CH2(S)+CO2=CO+CH2O

R287 OH+HO2=O2+H2O 42% O2 0.5 MPa


31% O2 0.5 MPa
R35 H+O2+H2O=HO2+H2O 42% O2 0.1 MPa
31% O2 0.1 MPa
R52 H+CH3(+M)=CH4(+M)

-0.1 0.0 0.1 0.3 0.4 0.5


Sensitivity coefficients

Fig. 12. Laminar burning velocity as a function of the adiabatic flame tem- R38 H+O2=O+OH (b)
perature.
R119 HO2+CH3=OH+CH3O

R99 OH+CO=H+CO2

R166 HCO+H2O=H+CO+H2O P = 0.1 MPa


R153 CH2(S)+CO2=CO+CH2O
= 1.0

R287 OH+HO2=O2+H2O 21% O2/(O2+N2)


31% O2/(O2+FCO2)
R35 H+O2+H2O=HO2+H2O 31% O2/(O2+N2)
31% O2/(O2+CO2)
R52 H+CH3(+M)=CH4(+M)

-0.1 0.0 0.1 0.3 0.4 0.5


Sensitivity coefficients

Fig. 14. Normalized sensitivity coefficients of the laminar burning velocity of


the 8 most rate limiting reactions in CH4/O2/CO2 flames for different O2 molar
fractions and pressures computed by the GRI-Mech 3.0. Top: 0.1 MPa; bottom:
O2 molar fraction = 0.38.

CO2 flames for different O2 molar fractions and pressures computed


Fig. 13. Laminar burning velocity as a function of the thermal diffusivity. using the GRI-Mech 3.0. As the CO2 molar fraction and pressure in-
crease, the sensitivity coefficients of almost all top reactions increase.
different SL values due to the different diffusivity properties. The most important chain-branching and chain-terminating reactions
Fig. 13 shows the laminar burning velocity as a function of the R38 (H + O2 = O + OH) and R52 [CH3 + H (+M) = CH4 (+M)] have
thermal diffusivity (α = λ/(ρCp), which is inversely proportional to the the largest positive and negative sensitivity coefficients for all condi-
pressure. As the pressure increases, both the SL and the thermal diffu- tions. Fig. 14(b) compares other three cases with the 31% O2/
sivity decrease. In fact, the Tad increases as the pressure increases re- (O2 + CO2) case at 0.1 MPa, and shows that CO2 participates in the
sulting in the increase of the total mass flow rate, while the upstream chemical reactions mainly through reactions R99
gas also becomes denser for the flame to heat up and pass through, (CO + OH = CO2 + H) and R153 [CH2(S) + CO2 = CO + CH2O],
leading to the decrease of the SL. The SL decrease rate is different for which are two important reactions that convert CO2 to CO. In addition,
each line, which means there are important factors influencing the la- the reverse of reaction R99 competes with reaction R38 in consuming H
minar burning velocity besides the thermal diffusivity or the density radicals retarding the chain-propagating reactions [8].
variation. The 31% O2/(O2 + CO2) and 31% O2/(O2 + FCO2) lines The active radicals H and OH play an important role on SL through
have the same thermal diffusivity, but different SL due to the chemical chain-propagating reactions [9,49]. Fig. 15 shows the radical profiles of
effect of the CO2, which retards the propagation of the flame. H and OH along with CO production for stoichiometric CH4/O2/CO2
A comparison between the 0.1 MPa and 0.5 MPa lines for the CH4/ flames in comparison with other three cases that have the same Tad,
O2/CO2 flames (Figs. 12 and 13), reveals that as the oxygen content thermal diffusivity and oxygen content at 0.1 and 0.5 MPa. As oxygen
increases, both the thermal diffusivity and the Tad increase lead to the content decreases and pressure increases, the mole fractions of the H,
increase of SL, while the increase rate of SL against Tad is greater than OH and CO decrease typifying the decrease of the overall reaction rate.
that against the thermal diffusivity. Hu et al. [45] and Chen et al. [46] It is seen that the mole fraction of OH is always higher than H, which is
reported that variations in Tad are the major cause for SL variations with more significant for CO2-diluted flames than N2-diluted and FCO2-di-
the addition of diluents, and that the thermal-diffusivity is of secondary luted flames. And H radical decreases much more than OH as pressure
importantce at 1 atm. At elevated pressures, both these two effects are increases for all cases leading to the decrease of SL. The active radicals
weakened and the effect of the Tad variation on SL is more significant in cases of the same oxygen mole fraction, the same Tad and the same
than that of the thermal diffusivity variation. thermal diffusivity as 31% O2/(O2 + CO2) are always higher than those
in the CO2-diluted flames, but CO2-diluted flame produces the highest
3.4. Sensitivity and kinetic analyses mole fraction of CO due to the reverse of reaction R99
(CO + OH = CO2 + H) and R153 [CH2(S) + CO2 = CO + CH2O], as
Fig. 14(a) shows the normalized sensitivity coefficients of the la- illustrated above.
minar burning velocity of the 8 most rate limiting reactions in CH4/O2/ To illustrate the chemical effect of CO2 on flame propagation,

7
S. Wang, et al. Fuel 259 (2020) 116152

0.015 0.0150
(a) (d) 0.5 MPa
0.1 MPa
0.0125 31% O2/(O2+CO2)
31% O2/(O2+CO2)
31% O2/(O2+FCO2)
31% O2/(O2+FCO2) 0.0100
0.010 31% O2/(O2+N2)
31% O2/(O2+N2)
21% O2/(O2+N2)
21% O2/(O2+N2)
0.0075

0.005 0.0050 H
H
0.0025

0.000 0.0000
(b) (e)
Species mole fraction

Species mole fraction


0.015 0.015

0.010 0.010
OH OH
0.005 0.005

0.000 0.000
(c) (f)
0.08 0.08

0.06 0.06

CO 0.04 CO
0.04

0.02 0.02

0.00 0.00
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
Distance (cm) Distance (cm)
Fig. 15. Mole fraction profiles of H, OH and CO for stoichiometric CH4/O2/diluents flames. Top: 0.1 MPa; bottom: 0.5 MPa.

Fig. 16 shows the reaction pathways for stoichiometric CH4/O2/FCO2 propagating flame and normalized by the total reaction rate [50].
and CH4/O2/CO2 flames at 0.1 and 0.5 MPa. The consumption of the Fig. 16 reveals that FCO2-diluted possesses similar reaction pathway as
active radical i (i = H, O and OH) through a reaction, ηi is computed by CO2-diluted, but the reactions that consume O and H to produce OH are
spatial integration (Si) within the reaction zone of the adiabatic freely- enhanced in the CO2-diluted flames like reactions (R38:

Fig. 16. Reaction pathways for stoichiometric CH4/O2/FCO2 (fictitious CO2) and CH4/O2/CO2 flames at 0.1 MPa. Data in brackets are for 0.5 MPa.

8
S. Wang, et al. Fuel 259 (2020) 116152

Fig. 17. Normalized [X]max-0.5MPa /[X]max-0.1MPa ratios of radicals H,OH and CH3 as a function of ϕ for CH4/O2/N2 and CH4/O2/CO2 flames (a, b); relative con-
tribution in consuming the radical H through reactions R38, ηH(R38) and R52, ηH(R52), as a function of ϕ for CH4/O2/N2 and CH4/O2/CO2 flames (c, d); normalized
SH − 0.5MPa S
S
(R38) ratio as a function of ϕ for CH4/O2/N2 and CH4/O2/CO2 flames (e, f); normalized SH − 0.5MPa (R52) ratio as a function of ϕ for CH4/O2/N2 and CH4/O2/
H − 0.1MPa H − 0.1MPa
CO2 flames (g, h).

H + O2 = O + OH) and R86 (O + H2O = 2OH) while the converse the definition of β and the chain-propagating role of the active radicals
reactions are weakened in the CO2-diluted flames like R3 on SL, Fig. 17a-b shows the normalized [X]max-0.5MPa/[X]max-0.1MPa ra-
(O + H2 = H + OH) and R99 (CO + OH = CO2 + H). As pressure in- tios of radicals H, OH and CH3 as a function of ϕ for CH4/O2/N2 and
creases, the enhanced three-body reactions like R33 CH4/O2/CO2 flames. These ratios represent the reduction extent of each
(H + O2 + M = HO2 + M) and R52 (H + CH3 + M = CH4 + M) lead active radical from 0.1 to 0.5 MPa. CH3 presents the slightest reduction
to the decrease of the amounts of active radicals, while the largest rate and H the largest one. For the three radicals, the ratios present a
enhancement is observed in reaction R86 (O + H2O = 2OH) - up to minimum at rich side, where the non-monotonic behavior is similar.
20% as pressure increases. Fig. 17c, d shows the relative contribution in consuming the radical
H through reactions R38, ηH(R38) and R52, ηH(R52), as a function of ϕ for
3.5. Non-monotonic behavior CH4/O2/N2 and CH4/O2/CO2 flames, which are the most important
chain-branching and chain-terminating reactions. As ϕ increases, the H
Figs. 7 and 10 revealed that β features a non-monotonic behavior radicals consumed through reaction R38 decrease, and conversely,
under fuel-rich flames for all conditons. The location of the turning more H radicals participate in reaction R52. This trend is intensified as
point shifts to the richer side as oxygen content increases. Considering the pressure increases since reaction R52 is a three-body collision

9
S. Wang, et al. Fuel 259 (2020) 116152

reaction. As ϕ further increases, more CH3 is consumed through reac- 2010;91:1617–23.


tion 2CH3 (+M) = C2H6 (+M), which slightly alleviates the decreasing [5] Park SK, Kim TS, Sohn JL, Lee YD. An integrated power generation system com-
bining solid oxide fuel cell and oxy-fuel combustion for high performance and CO2
and increasing trends of ηH-R38 and ηH-R52. capture. Appl Energy 2011;88:1187–96.
The competition of H through R38 and R52 plays a major role in the [6] de Persis S, Foucher F, Pillier L, Osorio V, Gökalp I. Effects of O2 enrichment and
S
variation of β. Fig. 17e–h shows the normalized SH − 0.5MPa (R38) and CO2 dilution on laminar methane flames. Energy 2013;55:1055–66.
H − 0.1MPa [7] Xie Y, Wang J, Zhang M, Gong J, Jin W, Huang Z. Experimental and numerical study
SH − 0.5MPa
SH − 0.1MPa
(R52) ratios as a function of ϕ for CH4/O2/N2 and CH4/O2/CO2 on laminar flame characteristics of methane oxy-fuel mixtures highly diluted with
CO2. Energy Fuels 2013;27:6231–7.
flames. These ratios represent the positive and negative effects on SL of
[8] Glarborg P, Bentzen LLB. Chemical effects of a high CO2 concentration in oxy-fuel
increasing the pressure. As the oxygen content increases, both combustion of methane. Energy Fuels 2008;22:291–6.
SH − 0.5MPa S
SH − 0.1MPa
(R38) and SH − 0.5MPa (R52) increase, and accordingly the overall [9] Wang Z, Wang S, Whiddon R, Han X, He Y, Cen K. Effect of hydrogen addition on
H − 0.1MPa
laminar burning velocity of CH4/DME mixtures by heat flux method and kinetic
reaction rate also increases, as happens with β in Figs. 7 and 10.
modeling. Fuel 2018;232:729–42.
Moreover, it is seen that the positive effect has a minimum at the rich [10] Wang ZH, Yang L, Li B, Li ZS, Sun ZW, Aldén M, et al. Investigation of combustion
side, where the negative effect has a maximum, indicating that the enhancement by ozone additive in CH4/air flames using direct laminar burning
velocity measurements and kinetic simulations. Combust Flame 2012;159:120–9.
overall reaction rate has a minimum, leading to the non-monotonic
[11] Dirrenberger P, Le Gall H, Bounaceur R, Herbinet O, Glaude P-A, Konnov A, et al.
behavior of the values of β. Measurements of laminar flame velocity for components of natural gas. Energy
Fuels 2011;25:3875–84.
4. Conclusions [12] Coppens FHV, De Ruyck J, Konnov AA. The effects of composition on burning ve-
locity and nitric oxide formation in laminar premixed flames of CH4 + H2 + O2 +
N2. Combust Flame 2007;149:409–17.
1. The heat flux method has been applied to measure strecthless [13] Hermanns RTE, Konnov AA, Bastiaans RJM, de Goey LPH, Lucka K, Kohne H.
adiabatic laminar burning velocity of CH4/O2/N2 flames up to Effects of temperature and composition on the laminar burning velocity of CH4 +
H2 + O2 + N2 flames. Fuel 2010;89:114–21.
0.5 MPa and CH4/O2/CO2 flames up to 0.3 MPa, with equivalence [14] Goswami M, Derks SCR, Coumans K, Slikker WJ, de Andrade Oliveira MH, Bastiaans
ratios vaying from 0.6 to 1.6 and oxygen molar fractions from 0.18 RJM, et al. The effect of elevated pressures on the laminar burning velocity of
to 0.42. Comparisons of the present SL data with literature data methane + air mixtures. Combust Flame 2013;160:1627–35.
[15] Goswami M, van Griensven JGH, Bastiaans RJM, Konnov AA, de Goey LPH.
showed good agreement at atmospheric and elevated pressures. Experimental and modeling study of the effect of elevated pressure on lean high-
Both the GRI-Mech 3.0 and HP-Mech mechanisms predicted rea- hydrogen syngas flames. Proc Combust Inst 2015;35:655–62.
sonably well the present results. [16] Dirrenberger P, Le Gall H, Bounaceur R, Glaude P-A, Battin-Leclerc F.
Measurements of laminar burning velocities above atmospheric pressure using the
2. Thermal-diffusion effects play a major role in the laminar burning
heat flux method—application to the case of n-pentane. Energy Fuels
velocity decrease due to CO2 dilution at normal and elevated pres- 2015;29:398–404.
sures. The adiabatic flame temperature has a dominant effect on the [17] Hu X, Yu Q, Liu J, Sun N. Investigation of laminar flame speeds of CH4/O2/CO2
mixtures at ordinary pressure and kinetic simulation. Energy 2014;70:626–34.
laminar burning velocity when the CO2 content varies and an op-
[18] Mazas AN, Lacoste DA, Schuller T. Experimental and numerical investigation on the
posite influence when the pressure increases. laminar flame speed of CH4/O2 mixtures diluted with CO2 and H2O. ASME. Turbo
3. Kinetic analysis indicated that the presence of active radicals de- expo: power for land, sea, and air, 2: combustion, fuels and emissions. 2010. p.
creases as oxygen molar fraction decreases and pressure increases 411–21.
[19] Oh J, Noh D. Laminar burning velocity of oxy-methane flames in atmospheric
leading to the decrease of the laminar burning velocity. The reverse condition. Energy 2012;45:669–75.
of CO + OH = CO2 + H leads to the increase of OH and CO mole [20] Konnov AA, Dyakov IV. Measurement of propagation speeds in adiabatic cellular
fractions and retards the flame propagation in competition with the premixed flames of CH4 + O2 + CO2. Exp Therm Fluid Sci 2005;29:901–7.
[21] Chen Z, Qin X, Xu B, Ju Y, Liu F. Studies of radiation absorption on flame speed and
reaction H + O2 = O + OH. flammability limit of CO2 diluted methane flames at elevated pressures. Proc
4. The pressure exponent β increases as the oxygen molar fraction in- Combust Inst 2007;31:2693–700.
creases. Competition of H consuming reaction H + O2 = O + OH [22] Khan AR, Anbusaravanan S, Kalathi L, Velamati R, Prathap C. Investigation of di-
lution effect with N2/CO2 on laminar burning velocity of premixed methane/
with CH3 consuming reactions 2CH3 (+M) = C2H6 and CH3 + H oxygen mixtures using freely expanding spherical flames. Fuel 2017;196:225–32.
(+M) = CH4 (+M) leads to a non-monotonic behavior of the [23] Cai X, Wang JH, Zhang WJ, Xie YL, Zhang M, Huang ZH. Effects of oxygen en-
overall reaction order for both N2- and CO2-diluted flames. richment on laminar burning velocities and Markstein lengths of CH4/O2/N2 flames
at elevated pressures. Fuel 2016;184:466–73.
[24] De Goey LPH, Van Maaren A, Quax RM. Stabilization of adiabatic premixed laminar
Acknowledgements flames on a flat flame burner. Combust Sci Technol 1993;92:201–7.
[25] Bosschaart KJ, de Goey LPH. Detailed analysis of the heat flux method for mea-
suring burning velocities. Combust Flame 2003;132:170–80.
This work was supported by the National Natural Science
[26] Han X, Wang Z, Wang S, Whiddon R, He Y, Lv Y, et al. Parametrization of the
Foundation of China (51621005) and the Program of Introducing temperature dependence of laminar burning velocity for methane and ethane
Talents of Discipline to University (B08026). The authors give special flames. Fuel 2019:1028–37.
thanks to the National Program for Support of Top-notch Young [27] Bosschaart KJ, de Goey LPH. The laminar burning velocity of flames propagating in
mixtures of hydrocarbons and air measured with the heat flux method. Combust
Professionals of China. M. Costa also acknowledges the support of Flame 2004;136:261–9.
Fundação para a Ciência e a Tecnologia, through IDMEC, under LAETA, [28] Sileghem L, Alekseev VA, Vancoillie J, Van Geem KM, Nilsson EJK, Verhelst S, et al.
project UID/EMS/50022/2013. Laminar burning velocity of gasoline and the gasoline surrogate components iso-
octane, n-heptane and toluene. Fuel 2013;112:355–65.
[29] Van Maaren A, Thung DS, De Goey LRH. Measurement of flame temperature and
Appendix A. Supplementary data adiabatic burning velocity of methane/air mixtures. Combust Sci Technol
1994;96:327–44.
[30] Alekseev VA, Naucler JD, Christensen M, Nilsson EJK, Volkov EN, de Goey LPH,
Supplementary data to this article can be found online at https:// et al. Experimental uncertainties of the heat flux method for measuring burning
doi.org/10.1016/j.fuel.2019.116152. velocities. Combust Sci Technol 2016;188:853–94.
[31] Smith GP, Golden DM, Frenklach M, Moriarty NW, Eiteneer B, Goldenberg M, et al.
GRI-Mech 3.0; 1999. http://www.me.berkeley.edu/gri_mech/.
References
[32] Yang Xueliang, Shen Xiaobo, Santer Jeffery, Zhao Hao, Ju Yiguang. A high pressure
mechanism for H2, CO, CH2O, CH4, CH3OH, C2H2, C2H4, C2H6 with EGR Effects
[1] Wall TF. Combustion processes for carbon capture. Proc Combust Inst (CO2 and H2O) and HCO prompt reactions. 2017. http://engine.princeton.edu/
2007;31:31–47. mechanism/HP-Mech.html.
[2] Lasek JA, Janusz M, Zuwała J, Głód K, Iluk A. Oxy-fuel combustion of selected solid [33] Rozenchan G, Zhu DL, Law CK, Tse SD. Outward propagation, burning velocities,
fuels under atmospheric and elevated pressures. Energy 2013;62:105–12. and chemical effects of methane flames up to 60 atm. Proc Combust Inst
[3] Toftegaard MB, Brix J, Jensen PA, Glarborg P, Jensen AD. Oxy-fuel combustion of 2002;29:1461–70.
solid fuels. Prog Energy Combust Sci 2010;36:581–625. [34] Tahtouh T, Halter F, Mounaïm-Rousselle C. Measurement of laminar burning speeds
[4] Czakiert T, Sztekler K, Karski S, Markiewicz D, Nowak W. Oxy-fuel circulating and Markstein lengths using a novel methodology. Combust Flame
fluidized bed combustion in a small pilot-scale test rig. Fuel Process Technol 2009;156:1735–43.

10
S. Wang, et al. Fuel 259 (2020) 116152

[35] Lowry W, de Vries J, Krejci M, Petersen E, Serinyel Z, Metcalfe W, et al. Laminar methanol kinetic mechanisms at oxy-fuel conditions. Combust Flame
flame speed measurements and modeling of pure alkanes and alkane blends at 2015;162:1719–28.
elevated pressures. J Eng Gas Turbines Power 2011;133:091501–9. [44] Bosschaart KJ, Versluis M, Knikker R, Vandermeer TH, Schreel KRAM, De Goey
[36] Gu XJ, Haq MZ, Lawes M, Woolley R. Laminar burning velocity and Markstein LPH, et al. The heat flux method for producing burner stabalized adiabatic flames:
lengths of methane–air mixtures. Combust Flame 2000;121:41–58. an evaluation with CARS thermometry. Combust Sci Technol 2001;169:69–87.
[37] Hassan MI, Aung KT, Faeth GM. Measured and predicted properties of laminar [45] Hu E, Jiang X, Huang Z, Iida N. Numerical study on the effects of diluents on the
premixed methane/air flames at various pressures. Combust Flame laminar burning velocity of methane-air mixtures. Energy Fuels 2012;26:4242–52.
1998;115:539–50. [46] Chen Z, Tang C, Fu J, Jiang X, Li Q, Wei L, et al. Experimental and numerical
[38] Egolfopoulos FN, Cho P, Law CK. Laminar flame speeds of methane-air mixtures investigation on diluted DME flames: thermal and chemical kinetic effects on la-
under reduced and elevated pressures. Combust Flame 1989;76:375–91. minar flame speeds. Fuel 2012;102:567–73.
[39] Egolfopoulos FN, Law CK. Chain mechanisms in the overall reaction orders in la- [47] Chen Z, Wei L, Huang Z, Miao H, Wang X, Jiang D. Measurement of laminar burning
minar flame propagation. Combust Flame 1990;80:7–16. velocities of dimethyl ether−air premixed mixtures with N2 and CO2 dilution.
[40] Konnov AA, Mohammad A, Kishore VR, Kim NI, Prathap C, Kumar S. A compre- Energy Fuels 2009;23:735–9.
hensive review of measurements and data analysis of laminar burning velocities for [48] Halter F, Foucher F, Landry L, Mounaïm-Rousselle C. Effect of dilution by nitrogen
various fuel+air mixtures. Prog Energy Combust Sci 2018;68:197–267. and/or carbon dioxide on methane and iso-octane air flames. Combust Sci Technol
[41] Halter F, Chauveau C, Djebaïli-Chaumeix N, Gökalp I. Characterization of the ef- 2009;181:813–27.
fects of pressure and hydrogen concentration on laminar burning velocities of [49] Han X, Wang Z, Costa M, Sun Z, He Y, Cen K. Experimental and kinetic modeling
methane–hydrogen–air mixtures. Proc Combust Inst 2005;30:201–8. study of laminar burning velocities of NH3/air, NH3/H2/air, NH3/CO/air and
[42] Chen Z. Effects of radiation absorption on spherical flame propagation and radia- NH3/CH4/air premixed flames. Combust Flame 2019;206:214–26.
tion-induced uncertainty in laminar flame speed measurement. Proc Combust Inst [50] Das AK, Kumar K, Sung C-J. Laminar flame speeds of moist syngas mixtures.
2017;36:1129–36. Combust Flame 2011;158:345–53.
[43] Naucler JD, Sileghem L, Nilsson EJK, Verhelst S, Konnov AA. Performance of

11

You might also like