Paper 1

You might also like

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

Acta Astronautica 134 (2017) 133–140

Contents lists available at ScienceDirect

Acta Astronautica
journal homepage: www.elsevier.com/locate/actaastro

Computational fluid dynamics analysis of a synthesis gas turbulent MARK


combustion in a round jet burner

Mohammad Mansourian, Reza Kamali
School of Mechanical Engineering, Shiraz University, Shiraz 71348-51154, Iran

A R T I C L E I N F O A BS T RAC T

Keywords: In this study, the RNG-Large Eddy Simulation (RNG-LES) methodology of a synthesis gas turbulent combustion
Jet burner in a round jet burner is investigated, using OpenFoam package. In this regard, the extended EDC extinction
RNG-LES model of Aminian et al. for coupling the reaction and turbulent flow along with various reaction kinetics
Skeletal and GRI-3.0 mechanisms mechanisms such as Skeletal and GRI-MECH 3.0 have been utilized. To estimate precision and error
Flow temperature
accumulation, we used the Smirinov's method and the results are compared with the available experimental
Inlet velocity
data under the same conditions. As a result, it was found that the GRI-3.0 reaction mechanism has the least
computational error and therefore, was considered as a reference reaction mechanism. Afterwards, we
investigated the influence of various working parameters including the inlet flow temperature and inlet velocity
on the behavior of combustion. The results show that the maximum burner temperature and pollutant emission
are affected by changing the inlet flow temperature and velocity.

1. Introduction Safer et al. [7] reported a capable model based on the combination of
transport equations and PDF (Probability Density Function) calcula-
The use of experimental methods for studying the combustion tions to study of a free axisymmetric turbulent jet and non-premixed
process as well as high speed of the flow is hampered due to the danger flames. Burke et al. [8] noted that the measured mass burning rate of
of structures destruction, therefore the development of numerical laminar premixed flames of H2 / CH4 /O2 / He mixture at equivalence
methods is of urgent interest. Computational fluid dynamics (CFD) ratios from 0.3 to 1.0 and pressures from 1 to 25 atm can be obtained
allow obtaining and visualizing detailed 3D flow patterns, identifying using the GRI Mech 3.0. However, some of the rate constants need to
both local features and integral properties of the processes [1–3]. be adjusted empirically. Kuznetsov et al. [9] concluded that the GRI-3.0
The H2 /O2 reaction system is a fundamental topic in combustion is reasonable to compute the laminar burning velocity of H2 /O2 /H2 O
science that has historically received significant attention due to both stoichiometric flame for pressures ranging from 10 to 72 bars, but the
its rich kinetic behavior and its importance to a variety of applications mechanism of Lutz [10] is better for 1–72 bars. The burning velocities
in energy conversion. Since H2 and the intermediate oxidation species calculated by Boushaki et al. [11] using the GRI-3.0 for CH4 / H2 / H2 O / air
are dominant reactants in the oxidation of all hydrocarbons and atmospheric flames over a wide range of equivalence ratios with 0–30%
oxygenated fuels, the H2 /O2 mechanism not only forms an essential H2 , 0% H2 O (dry) and 100% relative humidity (wet) conditions using
subset of any hydrocarbon or oxygenate oxidation mechanism [4], but the GRI-3.0, are in acceptable coincidence with measurements. He
also contains a number of reactions whose rate constants are among et al. [12] deduced that the laminar flame speeds calculated using the
the most sensitive for combustion predictions of all hydrocarbons and GRI Mech 3.0 and USC-II mechanisms agreed well with their
oxygenate fuels. In recent years, there has been an increased interest in measurements using PLIF techniques for lean flames; moreover,
studying the combustion of hydrogen, and fuel mixtures consisting of USC-II mechanism [13] gives better agreement for flames with an
carbon monoxide and hydrogen, referred to as “syngas” or “wet CO” equivalence ratio of 0.8 and 0.9. Vasu et al. [14] noted that the GRI-3.0
potentially including additional species such as CO2 and/or H2 O . is able to capture the trends and magnitudes of the measured ignition
Marzouk and Huckaby [5] compared the performance and computa- delays at temperatures 974–1160 K and pressures 1.1–2.6 atm for
tional cost of 8 kinetic models (3 global and 5 elementary) that describe stoichiometric H2 / CO2 / CO / air mixtures. The analysis of ignition data
the finite-rate chemistry of syngas combustion. They recommended the by Petersen et al. [15] for syngas/air at 600–1148 K and 10–30 atm
Westbrook-Dryer [6] model for modeling turbulent syngas flames. suggested that the available kinetic mechanisms are reasonable if the


Corresponding author.
E-mail address: rkamali@shirazu.ac.ir (R. Kamali).

http://dx.doi.org/10.1016/j.actaastro.2017.01.038
Received 29 October 2016; Accepted 29 January 2017
Available online 07 February 2017
0094-5765/ © 2017 Published by Elsevier Ltd on behalf of IAA.
M. Mansourian, R. Kamali Acta Astronautica 134 (2017) 133–140

Nomenclature Yk Mass fraction of kth

H Enthalpy Greek letters


k Order of accuracy of numerical schemes
N1 Number of cells in the direction of integration δ Kronecker delta
P Pressure εSGS Subgrid scale dissipation
Prt Prantl number (unity) µSGS Subgrid scale viscosity
u Velocity vector, ρ Density
S max Total error τ Stress tensor
Sct Schmidt number (unity)
Sij Grid scale rate of stress tensor

temperature is larger than 1000 K, even up to 450 atm as shown in conditions. In this regards, RNG-LES model in a finite volume frame-
[16]. Two points become clear from this brief survey; the BFG like fuel work, two different reaction mechanisms such as Skeletal [24,25] and
mixture has not been considered in earlier studies and the use of the GRI-3.0 [13] as well as the extended EDC extinction model of Aminian
GRI Mech 3.0 is reasonable as long as the fuel mixture contains et al. [26] to well-resolve turbulence–chemistry interaction field and
molecules such as H2,CO , CH4 , CO2 and H2 O . also to estimate precision and error accumulation we used the
Bhargava et al. [17] investigated an industrial gas turbine fueled Smirinov's method [27–31] have utilized in openfoam package. The
with natural gas under 0.1–0.4 MPa pressure through measurements results are compared with the available experimental data [32,33]
and perfectly stirred reactor network simulation. They found that the under the same conditions. Finally, the effects of inlet temperature and
NOx pressure exponent varies from −0.18 to 1.6 for the tangential velocity have investigated on the combustion behavior in the round jet
nozzle and −0.77 to 0.61 for the axial nozzle as the equivalence ratio burner.
increases from 0.43 to 0.65. Moreover, when the equivalence ratio
range extends, the NOx exponent variation might not be monotonous. 2. Governing equations
Göke et al. [18] investigated NOx and CO emissions in a premixed
methane flame by measurements and simulation within pressure range The naturally three dimensional and unsteady turbulent flows are
of 1.5–9.0 bar. They found that the NOx pressure exponent ranges from often influenced by strong non-homogeneous effects and rapid trans-
0.1 to 0.65, and the CO pressure exponent is about −0.4 at low formations which prevent using the isotropic models in simulations
temperature and about −0.5 at higher temperature. Smirnov et al. [19] [34,35]. The modified sub-grids methods which can improve LES
studied numerical simulation of detonation engine fed by fuel-oxygen methodologies are also presented, although their usage is confined due
mixture, using simplified kinetics of acetylene burning includes 11 to their natural complexity. Therefore, new methods for simulating the
reactions with 9 components. The advantage of a constant volume subgrid scales are in progress. In common methods some models are
combustion cycle as compared to constant pressure combustion was in presented to consider the effects of subgrid scales in filtered Navier-
terms of thermodynamic efficiency focused for advanced propulsion on Stokes equations [36]. LES is more common than the other models. In
detonation engine. this model at first large scales should be solved with energy which is
In the work of Sigfrid et al. [20], a natural gas combustor was given to them. For this purpose, the effect of the small scales which are
simulated by the chemical reactor network model under pressure range waste, should be considered in energy balance. In LES methodology,
of 3–9 bar, and the pathways of NOx formation was analyzed at the filter is used to separate the large scales from small ones [37].
elevated pressures and different equivalence ratios. They found that Therefore, all variables like f are divided into two parts, grid and sub-
∼ ∼
the higher pressure enhances the N2 O pathway under all conditions, grid. In other words, f = f + f ′ where f = ρ f / ρ∼ is sub-grid tension
though thermal pathway might still be the most active at higher [37].
temperature. For syngas, rare CH radical was produced in the By filtering the variables in the compressible Navier-Stokes equa-
combustion process, and the prompt and NNH pathways could be tions, the instantaneous filtered equation leads to the following
considered the same [21]. It implied that the mixedness level, which equations. A more detailed discussion can be found in the textbooks
would vary with combustion mode and nozzle form, has effects on by Poinsot and Veynante [38].
emissions. For the CH4− air combustor, Han et al. [22] found that the
∂ρ ∂ ∼
NOx emissions scaled with pressure aspen, where "pressure exponent + (ρui ) = 0
∂t ∂xi (1)
for NOx "=0.5 under premix combustion, and "pressure exponent of
∂ ⎡⎛ μ ⎞ ∂Y∼ ⎤
NOx "=1.1 for fuel staged combustion. Biagioli and Gu the [23] ∼
∂ρ Yk ∂ ∼∼
numerically studied the effects of fuel–air unmixedness on the NOx + (ρui Yk ) = ⎢ ⎜μ + SGS ⎟ k ⎥ + ω̇ k = 1, ... , N
∂t ∂xi ∂xi ⎣ ⎝ Sct ⎠ ∂xi ⎦ (2)
emissions of industrial premixed gas turbine burners fueled with
natural gas from 1 to 30 bars. It could be concluded that the
∂ρ u∼i ∂ ∼∼ ∂p ∂ ⎡⎢ ⎛ ∂u∼ ∂u∼ ⎞ δij ⎤⎥
unmixedness increases NOx emissions and the higher the pressure, + (ρui uj ) + = τij + μSGS ⎜ i + i ⎟ + Tkk
the greater the effect of unmixedness on NOx emissions. For syngas, ∂t ∂xi ∂xi ∂xi ⎢⎣ ⎝ ∂xj ∂xi ⎠ 3 ⎥⎦ (3)
diffusion combustion is widely adopted in commercial gas turbines to
∂ ⎡⎛ μ ∂h ⎤
∼ ∼
avoid flashback due to the higher burning speed. However, the ∂ρ h ∂ ∼∼ Dp μ ⎞
+ (ρui h ) = + ⎢⎜ + SGS ⎟ + ⎥
diffusion burning mode may not be beneficial for emission reducing ∂t ∂xi Dt ∂xi ⎣ ⎝ Pr Prt ⎠ ∂xi ⎦ (4)
compared with the premixed mode, which makes premixed syngas
In the above equations, the symbols ˜ and ¯ indicate filtered Faver
combustor a future research direction. Emissions during syngas
and filtered quantities, respectively, and the isentropic contribution,
combustion by the traditional burning mode are compared with the
Tkk, in equation (which is twice the sub-grid scale turbulent kinetic
advanced mode in this work.
energy) is unknown and usually absorbed into the filtered pressure. All
In our work, the behavior of syngas combustion is analyzed within
terms in the above equations are closed except the species transport's
wide temperature range and various inlet mass flow rate which may
(Eq. (2)) source term. Furthermore, the sub-grid scale kinetic energy
help to explain the emission behaviors of the combustor under different
obtained by,

134
M. Mansourian, R. Kamali Acta Astronautica 134 (2017) 133–140

∂KSGS ∂u∼j KSGS ∂ ⎛ μSGS ∂KSGS ⎞ ∂u∼j of time steps used to obtain the result:
+ = ⎜ ⎟ − τij − εSGS
∂t ∂xi ∂xi ⎝ ρ Prt ∂xi ⎠ ∂xi (5) RS = n max / n (16)
In which, The Eq. (16) characterizes reliability of results, i.e. how far below
Ce KSGS KSGS the limit the simulations were finalized. Indirectly it characterizes the
εSGS = accumulated error. The higher is the value of Eq. (16), the lower is the
Δ (6)
error. On tending Rs to unity the error tends to maximal allowable
μSGS 2 value.
τij = −2 Sij + KSGS δij
ρ 3 (7)
where, Ck and Ce are set to 0.2 and 1.048 in this study. In addition the 4. Raction mechanisms
grid scale rate of stress tensor (Sij ) is defined as follows:
4.1. Skeletal reduction methods
1 ⎛ ∂u∼ ∂u∼j ⎞
Sij = ⎜ i + ⎟
2 ⎝ ∂xj ∂xi ⎠ (8) A skeletal mechanism is generated from the detailed model by
excluding all elementary reactions in which the excluded species
Furthermore, the µSGS is the sub-grid scale viscosity which based on appear as reacting partners. Inert gases, such as Ar, He, and N2, will
Renormalization Group Theory: appear as chemically insignificant, but these are needed in the
μsgs = μ [1 + H (x )]1/3 mechanism for temperature and collision effects. Some small radicals,
(9)
such as CH*, CH, and OH*, are useful as signatures for flame surface.
The function H(x) is defined as: Recently, this mechanism is increased research activities focused on
⎧ x; x > 0 natural gas, syngas, and biogas. Detailed mechanisms for these fuels
H (x ) = ⎨ usually involve a smaller number of species that can be identified using
⎩ 0; x ≤ 0 (10)
species sensitivity methods. The skeletal mechanism is suitable for
x is given by x = μs μsgs / μ3 − C , where C=100 and μs is given by: equivalence ration of 0.5–15, pressure of 1–15 atm as well as
temperature of higher than 600 K. The skeletal mechanisms presented
μs = ρ [Crng Vol1/3]2 S (11) in this work include 31 species and 188 reactions [24,25].
with Crng=0.157. Furthermore, a finite volume framework with a pure
second order central differencing scheme and a blended scheme Linear 4.2. GRI-Mech 3.0 raction mechanism
Upwind Stabilized Transport [39] are considered in this study. In
addition, the boundary conditions for turbulent characteristics are GRI-Mech 3.0 is an optimized mechanism designed to model
tabulated in Table 1. natural gas combustion, including no formation and reburn chemistry
and this reaction mechanism is a compilation of 325 elementary
3. Precision estimates chemical reactions and associated rate coefficient expressions and
thermochemical parameters for the 53 species involved in them. The
Estimating precision and error accumulation is necessary for large conditions for which GRI-Mech 3.0 was optimized, limited primarily by
scale simulations of complex gas dynamics in unsteady state flows availability of reliable optimization targets, are roughly temperature
[27,28]. Depending on spatial resolution and numerical solver used a from 1000 to 2500 K, pressure of lower than 10 atm, and equivalence
definite error occurs in integration at each time step. This error can be ratio from 0.1 to 5 for premixed systems. Furthermore, GRI-Mech has
expressed in absolute or relative values. Accumulation of error takes been optimized for methane and natural gas as fuel. As such it does
place for successive time steps. To this end, the following formulas are include reactions that are involved in the combustion of other hydro-
used that have been proposed by Smirinov et al. [27,28]. carbon constituents of natural gas [13].

⎛ ΔL ⎞k +1
S1 ≈ ⎜ ⎟ 5. Geometry and boundary condition
⎝ L1 ⎠ (12)
⎛ 1 ⎞k +1 The burner studied in this model consists of a straight nozzle placed
which in uniform grid it could be S1=⎜ N ⎟ . In the presence of several in a slight co-flow (see Fig. 1). The model set-up corresponds to the one
⎝ 1⎠
directions in integration the errors are being summed up [27,28]: studied by Couci et al. [32] and the temperature and composition
resulting from the non-premixed combustion in the burner setup have
3
also been experimentally investigated by Barlow and co-workers [33].
Serr = ∑ Si
i =1 (13) As it is evident from the Fig. 1, the symmetry boundary condition is
used for the center plane while velocity inlet condition is chosen for
The allowable value of total error after the solution is finalized
S max
inlet nozzle. Full details of the geometry is presented in Table 2.
should be defined by the user of the simulator. However, we presume it
The fuel gas consists of three species; CO, H2 and N2 which is fed
should be between 1% and 5%, because initial and boundary conditions
are usually not known with a higher accuracy. Then the following
Table 1
inequality should be satisfied [27,28]: boundary conditions for turbulent characteristics.
Serr⋅ n ≤ S max (14)
Name Location Value Unit
Where n is the number of time steps in Naviere Stokes equations
kinetic energy Inlet 0.128 m2/s2
integration as well as the maximum allowable number of time steps for
kinetic energy Intial value in computational 0.128 m2/s2
solving a problem for each code could be determined by [27,28]: domain

n max = (S max / Serr )2 (15) Turbulent dissipation Inlet 0.49 m2/s3


rate
There can be introduced another important characteristic for the
Turbulent dissipation Intial value in computational 13 m2/s3
results of supercomputer simulations. That is the ratio of the maximal rate domain
allowable number of time steps for the problem, and the actual number

135
M. Mansourian, R. Kamali Acta Astronautica 134 (2017) 133–140

through the nozzle using an inlet velocity and temperature of 76 m/s


and 292 K respectively, while the co-flow (air) velocity outside of the
nozzle is 0.7 m/s. Furthermore, the Reynolds number for the jet is
approximately 16,700, based on the inlet velocity and the inner
diameter of the nozzle, indicating that the jet is fully turbulent. Full
details of the boundary conditions are indicated in Table 3. At the
nozzle exit, the fuel gas mixes with the co-flow, creating an unconfined
circular jet. A continuous reaction requires that the reactants and the
oxidizer be mixed at stoichiometric conditions. In this set-up, the
turbulent flow of the jet will effectively mix the fuel from the nozzle
with the co-flowing oxygen. In this burner the small recirculation zones
generated by the nozzle wall thickness provide the means to decelerate
hot product gas. Furthermore, the turbulent mixing in the outer parts
of the jet acts to accelerate fluid originating in the co-flow, and
incorporate it in the jet. This is commonly referred to as entrainment
and can be observed in the co-flow streamlines which bend towards the
jet downstream of the orifice.
The mass transport in the reacting jet is modeled by solving the
mass fractions of 53 species for GRI-3.0 mechanism and 31 species for
Skeletal mechanism. Under these circumstances, both the mixing and
the reaction processes in the jet are significantly influenced by the
turbulent nature of the flow. To account for the turbulence when
Fig. 1. Schematic of burner configuration. solving the flow field, the RNG-LES turbulence model is applied.
Taking advantage of the symmetry condition, half of geometry is
modeled three dimensionally using a cylindrical coordinate system.
Table 2
Full details of the geometry. Fig. 2 shows a schematic of grid structure that is used in this study.
The domain consists of 1.2e7 structured computational cells after
name Expression Value studying the results of grid independency. This observation is due to
inlet fuel and oxidizer, furthermore the grid is more dense around the
Pipe inner diameters (Di), mm – 4.58
Pipe thickness, mm – 0.88 inlet nozzle.

Pipe length (Pl), mm 10Di 45.8


6. Results and discussion
Geometry width, mm 50Di 229
Geometry height, mm Pl+120Di 595.4
6.1. Validation

Table 3 In this section, the validation of the present numerical method has
Full details of the boundary conditions. been investigated. To achieve this purpose, the numerical predictions
of the mass fraction of CO,CO2 , H2 O,H2, N2 , O2 as well as normalized
name Value unit
temperature and velocity distributions are shown in Fig. 3. In order to
Jet velocity 76 m s−1 compare various reaction mechanisms and experimental data [32,33].
Coflow velocity 0.7 m s−1 The results indicate that the GRI-3.0 mechanism is more accurate than
Skeletal mechanism in order to predict all combustion parameters.
Inlet temperature 292 K
Inlet volume fraction of CO 0.4 –
This is maybe because of the fact that the GRI-3.0 mechanism is able to
Inlet volume fraction of H2 0.3 – understand the local extinction; therefore, concerning the high accu-
Inlet volume fraction of N2 0.3 – racy of the GRI-3.0 mechanism results, this is considered as the basis
Inlet volume fraction of O2 (coflow) 0.21 – for all investigations available in the present research.
Inlet volume fraction of N2 (coflow) 0.79 –
After validating the numerical method, the effect of changing the
inlet velocity as well as the inlet temperature on the combustion

Fig. 2. Configuration of 3D cells at different locations of the computational domain.

136
M. Mansourian, R. Kamali Acta Astronautica 134 (2017) 133–140

Fig. 3. Comparison between current numerical results and experimental data [32,33].

Fig. 4. Influence of variation in velocity inlet on various combustion behaviors.

137
M. Mansourian, R. Kamali Acta Astronautica 134 (2017) 133–140

Fig. 5. Influence of variation in temperature inlet on various combustion parameters.

behavior is investigated. The results are described in the following temperature [40] which is shown in Fig. 4. The distribution of
sections. temperature in the fuel and oxidizer nozzle demonstrates that the
maximum temperature is fairly decreasing as the inlet velocity
increases, which could be the reason of the appearance of incomplete
6.2. Effect of inlet velocity combustion. This is consistent with the numerical results of Mardani
et al. [41], Mousavi and Abolfazli-Esfahani [42] and also the experi-
In this section, the influences of changing the inlet velocity (30, 76 mental results of Medwell et al. [43]. Furthermore, it is shown that with
and 100 m s−1) on the combustion process are studied in terms of inlet velocity increasing, the amount of CO2 production at first
complete or incomplete reaction, pollutant emissions as well as the increases and when it reaches its production maximum value its
combustion chamber temperature and the results are shown in Fig. 4. amount starts to decrease. Therefore, inlet velocity increasing does
The figure shows the distribution of CO,CO2 , H2 O,H2, N2 and normal- not have an effect on average CO2 production. In addition, the
ized temperatures at centerline and z/d of 20 and 50. It is shown that distribution of hydrogen and oxygen in combustion chamber length
when the inlet velocity is increasing, the measure of CO is also for different speeds entry of fuel and oxidizer show that the rise in the
increased, which shows the bad effects of the inlet velocity increase. inlet velocity increases the amount of hydrogen and oxygen mass
The main reason to increase of CO emission is a reduction in

138
M. Mansourian, R. Kamali Acta Astronautica 134 (2017) 133–140

fraction. It shows the great difference with complete combustion due to tion, Prog. Energy Combust. Sci. 10 (1984) 1–57.
[5] O.A. Marzouk, E.D. Huckaby, A Comparative study of eight finite-rate chemistry
changes in the stoichiometric process as well as reduction in the flow kinetics for Co/H2 combustion, Eng. Appl. Comput. Fluid Mech. 4 (2010) 331–356.
residence time. Increase and decrease in the amount of water vapor in [6] C.K. Westbrook, F.L. Dryer, Simplified reaction mechanisms for the oxidation of
the combustion chambers indicates approaching complete process and hydrocarbon fuels in flames, Combust. Sci. Technol. 27 (1981) 31–43.
[7] K. Safer, A. Bounif, M. Safer, I. Gökalp, Free turbulent reacting jet simulation based
getting far from complete combustion process respectively. According on combination of transport equations and PDF, Eng. Appl. Comput. Fluid Mech. 4
to Fig. 4, it is clear that the increase in the inlet flow velocity in parts of (2010) 246–259.
the chamber, increases the produced water vapor and in some parts [8] M.P. Burke, F.L. Dryer, Y. Ju, Assessment of kinetic modeling for lean H2/CH4/O2/
diluent flames at high pressures, Proc. Combust. Inst. 33 (2011) 905–912.
decreases produced water vapor. This is caused by the fact that an [9] M. Kuznetsov, R. Redlinger, W. Breitung, J. Grune, A. Friedrich, N. Ichikawa,
increase in the inlet velocity decreases the necessary time for complete Laminar burning velocities of hydrogen-oxygen-steam mixtures at elevated tem-
reaction of fuel and oxidizer. peratures and pressures, Proc. Combust. Inst. 33 (2011) 895–903.
[10] A.E. Lutz, A numerical study of thermal ignition, in,; Sandia National Labs.,
Livermore, CA (USA), pp. Medium: X; Size: Pages: 112, 1988.
6.3. Effect of inlet temperature [11] T. Boushaki, Y. Dhué, L. Selle, B. Ferret, T. Poinsot, Effects of hydrogen and steam
addition on laminar burning velocity of methane–air premixed flame: experimental
In this section, the effects of changes in fuel and oxidizer inlet and numerical analysis, Int. J. Hydrog. Energy 37 (2012) 9412–9422.
[12] Y. He, Z. Wang, L. Yang, R. Whiddon, Z. Li, J. Zhou, K. Cen, Investigation of
temperature on combustion process behavior are investigated. To laminar flame speeds of typical syngas using laser based Bunsen method and
achieve this purpose, five various inlet temperatures are chosen that kinetic simulation, Fuel 95 (2012) 206–213.
are lower than self-ignition temperature of methane-air combustion [13] G.P. Smith, D.M. Golden, M. Frenklach, N.W. Moriarty, B. Eiteneer, M.
Goldenberg, C.T. Bowman, R.K. Hanson, S. Song, W.C. Gardiner, Jr., V.V.
such as 200, 300, 400, 500 and 600 K. Fig. 5 shows distributions of Lissianski, Z. Qin, 〈http://www.me.berkeley.edu/gri_mech/〉.
CO,CO2 , H2, N2 , O2 mass fractions in Z-direction as well as centerline [14] S.S. Vasu, D.F. Davidson, R.K. Hanson, Shock tube study of syngas ignition in rich
temperature, normalized velocity at z/d of 20 and 50 for different CO2 mixtures and determination of the rate of H+O2+CO2 → HO2+CO2, Energy
Fuels 25 (2011) 990–997.
temperatures. As shown in Fig. 5, the amount of CO,H2, N2 , O2 mass [15] E.L. Petersen, D.M. Kalitan, A.B. Barrett, S.C. Reehal, J.D. Mertens, D.J. Beerer,
fractions decrease as the inlet temperature increases, which indicate R.L. Hack, V.G. McDonell, New syngas/air ignition data at lower temperature and
that the combustion is completed. Reduction in H2 mass fraction, as an elevated pressure and comparison to current kinetics models, Combust. Flame 149
(2007) 244–247.
original fuel in the intended combustion process, shows that unburned
[16] R. Sivaramakrishnan, A. Comandini, R.S. Tranter, K. Brezinsky, S.G. Davis,
H2 is decreased and accordingly, the combustion is complete and more H. Wang, Combustion of CO/H2 mixtures at elevated pressures, Proc. Combust.
fuel is incorporated in process. Furthermore, the mass fraction of CO2 Inst. 31 (2007) 429–437.
[17] A. Bhargava, D.W. Kendrick, M.B. Colket, W.A. Sowa, K.H. Casleton, D.J. Maloney,
grows by increasing the inlet temperature. As mentioned in the above
Pressure effect on NOx and CO emissions in industrial gas turbines, in: ASME
discussion, more complete combustion is achieved by increasing the Turbo Expo 2000: Power for Land, Sea, and Air, American Society of Mechanical
inlet temperature. The growth of the CO2 mass fractions shows the Engineers, 2000, pp. V002T002A017-V002T002A017.
validity of this claim. In addition, the normalized temperature and [18] S. Göke, S. Schimek, S. Terhaar, T. Reichel, K. Göckeler, O. Krüger, J. Fleck,
P. Griebel, C.O. Paschereit, Influence of pressure and steam dilution on NOx and
velocity decrease as the inlet temperature increases. CO emissions in a premixed natural gas flame, J. Eng. Gas Turbines Power 136
(2014) 091508.
7. Conclusion [19] N.N. Smirnov, V.B. Betelin, V.F. Nikitin, Y.G. Phylippov, J. Koo, Detonation engine
fed by acetylene–oxygen mixture, Acta Astronaut. 104 (2014) 134–146.
[20] I.R. Sigfrid, R. Whiddon, R. Collin, J. Klingmann, Experimental and Reactor
In the present study the RNG-LES methodology in the Openfoam Network Study of Nitrogen Dilution Effects on NOX Formation for Natural Gas and
package is used to investigate the behavior of a syngas round jet Syngas at Elevated Pressures, in: ASME Turbo Expo 2013: Turbine Technical
Conference and Exposition, American Society of Mechanical Engineers, pp.
burner. The simulation is conducted with various reaction kinetics V01AT04A023-V001AT004A023, 2013.
mechanisms such as Skeletal and GRI MECH-3 and the results are [21] T.C. Lieuwen, V. Yang, R.A. Yetter, Synthesis Gas Combustion: Fundamentals and
compared with the available experimental data [32,33] under the same Applications, CRC Press, Boca Raton, 2010.
[22] D.-S. Han, G.-B. Kim, H.-S. Kim, C.-H. Jeon, Experimental study of NOx
conditions. Consequently, it was found that the GRI-3 reaction correlation for fuel staged combustion using lab-scale gas turbine combustor at
mechanism has the slightest computational error with a higher high pressure, Exp. Therm. Fluid Sci. 58 (2014) 62–69.
accuracy than skeletal mechanism by 3% (on average), therefore GRI [23] F. Biagioli, F. Güthe, Effect of pressure and fuel–air unmixedness on NOx
emissions from industrial gas turbine burners, Combust. Flame 151 (2007)
MECH-3 was considered as the reference reaction mechanism.
274–288.
Following that, the results of changes in inlet velocity show that the [24] S.G. Davis, A.V. Joshi, H. Wang, F. Egolfopoulos, An optimized kinetic model of
mass fraction of CO increases as the inlet velocity grows. Furthermore, H2/CO combustion, Proc. Combust. Inst. 30 (2005) 1283–1292.
changes in the inlet velocity do not have an effect on the average [25] B. Akih-Kumgeh, J.M. Bergthorson, Skeletal chemical Kinetic mechanisms for
syngas, methyl Butanoate, n-heptane, and n-decane, Energy Fuels 27 (2013)
CO2 production, and the maximum temperature is fairly decreasing as 2316–2326.
the inlet velocity increases. It should be noted that by increasing the [26] J. Aminian, C. Galletti, L. Tognotti, Extended EDC local extinction model
velocity from 30 to 76 m s−1, the changes in behavior of combustion are accounting finite-rate chemistry for MILD combustion, Fuel 165 (2016) 123–133.
[27] N.N. Smirnov, V.B. Betelin, R.M. Shagaliev, V.F. Nikitin, I.M. Belyakov,
more noticeable than the case with velocity increase of 76–100 m s−1. Y.N. Deryuguin, S.V. Aksenov, D.A. Korchazhkin, Hydrogen fuel rocket engines
In addition, the effect of inlet temperature was investigated, which is simulation using LOGOS code, Int. J. Hydrog. Energy 39 (2014) 10748–10756.
reflected in the decrease of the amounts of CO,H2, N2 , O2 mass fractions [28] N.N. Smirnov, V.B. Betelin, V.F. Nikitin, L.I. Stamov, D.I. Altoukhov, Accumulation
of errors in numerical simulations of chemically reacting gas dynamics, Acta
by up to 8% as the inlet temperature increases, meaning the combus- Astronaut. 117 (2015) 338–355.
tion is completed. Furthermore, the mass fraction of CO2 grows up to [29] R. Kamali, S.M. Mousavi, A.R. Binesh, Three dimensional CFD investigation of
8% by increasing the inlet temperature. shock train structure in a supersonic nozzle, Acta Astronaut. 116 (2015) 56–67.
[30] E. Goshtasbi Rad, S.M. Mousavi, Wall modeled large eddy simulation of supersonic
flow physics over compression–expansion ramp, Acta Astronaut. 117 (2015)
References 197–208.
[31] R. Kamali, S.M. Mousavi, D. Khojasteh, Three-dimensional passive and active
control methods of shock wave train physics in a duct, Int. J. Appl. Mech. 08 (2016)
[1] V.B. Betelin, N.N. Smirnov, V.F. Nikitin, V.R. Dushin, A.G. Kushnirenko,
1650047.
V.A. Nerchenko, Evaporation and ignition of droplets in combustion chambers
[32] A. Cuoci, A. Frassoldati, G. Buzzi Ferraris, T. Faravelli, E. Ranzi, The ignition,
modeling and simulation, Acta Astronaut. 70 (2012) 23–35.
combustion and flame structure of carbon monoxide/hydrogen mixtures. Note 2:
[2] N.N. Smirnov, V.B. Betelin, A.G. Kushnirenko, V.F. Nikitin, V.R. Dushin,
fluid dynamics and kinetic aspects of syngas combustion, Int. J. Hydrog. Energy 32
V.A. Nerchenko, Ignition of fuel sprays by shock wave mathematical modeling and
(2007) 3486–3500.
numerical simulation, Acta Astronaut. 87 (2013) 14–29.
[33] R.S. Barlow, G.J. Fiechtner, C.D. Carter, J.Y. Chen, Experiments on the scalar
[3] N.N. Smirnov, V.F. Nikitin, V.R. Dushin, Y.G. Filippov, V.A. Nerchenko, J. Khadem,
structure of turbulent CO/H2/N2 jet flames, Combust. Flame 120 (2000) 549–569.
Combustion onset in non-uniform dispersed mixtures, Acta Astronaut. 115 (2015)
[34] R. Kamali, S.M. Mousavi, A.R. Binesh, J. Abolfazli-Esfahani, Large eddy simulation
94–101.
of the flameless oxidation in the IFRF furnace with varying inlet conditions, Int. J.
[4] C.K. Westbrook, F.L. Dryer, Chemical kinetic modeling of hydrocarbon combus-

139
M. Mansourian, R. Kamali Acta Astronautica 134 (2017) 133–140

Spray. Combust. Dyn. (2016) 1–14. [40] D.T. Gottuk, R.J. Roby, C.L. Beyler, The role of temperature on carbon monoxide
[35] S.M. Mousavi, E. Roohi, Large eddy simulation of shock train in a convergent– production in compartment fires, Fire Saf. J. 24 (1995) 315–331.
divergent nozzle, Int. J. Mod. Phys. C 25 (2013) 1450003. [41] A. Mardani, S. Tabejamaat, M. Ghamari, Numerical study of influence of molecular
[36] K. Won-Wook, M. Suresh, A new dynamic one-equation subgrid-scale model for diffusion in the Mild combustion regime, Combust. Theory Model. 14 (2010)
large eddy simulations, in: Proceedings of the 33rd Aerospace Sciences Meeting 747–774.
and Exhibit, American Institute of Aeronautics and Astronautics, 1995. [42] S.M. Mousavi, J. Abolfazli-Esfahani, Numerical investigation of the flameless
[37] M. Chapuis, C. Fureby, E. Fedina, N. Alin, J. Tegnér, Les Modeling of Combustion oxidation of natural gas in the IFRF furnace using large eddy simulation, Int. J.
Applications using OpenFOAM, ECCOMAS CFD, 2010, pp. 14–17. Spray. Combust. Dyn. 6 (2014) 387–410.
[38] T. Poinsot, D. Veynante, Theoretical and Numerical Combustion, Edwards, [43] P.R. Medwell, P.A.M. Kalt, B.B. Dally, Simultaneous imaging of OH, formaldehyde,
Philadelphia, 2005. and temperature of turbulent nonpremixed jet flames in a heated and diluted
[39] H. Weller, Controlling the computational modes of the arbitrarily structured C grid, coflow, Combust. Flame 148 (2007) 48–61.
Mon. Weather Rev. 140 (2012) 3220–3234.

140

You might also like