Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

International Journal of Hydrogen Energy 30 (2005) 1113 – 1126

www.elsevier.com/locate/ijhydene

The effect of thermal radiation and radiation models on


hydrogen–hydrocarbon combustion modelling
Mustafa Ilbas∗
Department of Mechanical Engineering, School of Engineering, Erciyes University, 38039 Kayseri, Turkey

Received 26 July 2004; received in revised form 1 September 2004


Available online 11 November 2004

Abstract
This paper presents numerical simulation results from the modelling of a turbulent non-premixed hydrogen (H2 ) and
hydrogen–hydrocarbon flame with and without radiation models. CFD studies using Fluent code were carried out for three
cases; without radiation model, with the P-1 radiation model and with the discrete transfer radiation model. The model results
from these three cases are compared with each other and with the experimental results. The effects of fuel composition from
pure natural gas to hydrogen (100% CH4 , 70% H2 + 30% CH4 , and 100% H2 ) were also investigated. The predictions
are validated and compared against the experimental results obtained in this study and results from the literature. Turbulent
diffusion flames are investigated numerically using a finite volume method for the solution of the conservation equations and
reaction equations governing the problem. The standard k– model was used for the modelling of the turbulence phenomena
in the combustor. The chemical combustion reactions are described by seven species and three steps. A NOx post-processor
has been used for predicting NOx emissions from the combustor. Using the both radiation models caused over all lower
temperature levels for all fuel compositions compared with the results obtained using no radiation model. The results with the
radiation models are in better agreement with the measurements compared with the results without radiation model. The use
of a radiation model predicts lower over all temperature, thus lower NOx emissions.
䉷 2004 International Association for Hydrogen Energy. Published by Elsevier Ltd. All rights reserved.
Keywords: Combustion modelling; Hydrogen and hydrocarbon combustion; Radiation models; NOx formation

1. Introduction or hydrogen–hydrocarbon composite fuels is a great inter-


est to researchers and designers because of the complexity
Thermal radiation heat transfer involving gases can be of combustion process and flow field in such devices. Ig-
an important mode of heat transfer in high-temperature noring thermal radiation heat transfer may cause significant
combustors and furnaces, even under non-soot conditions. errors in the overall predictions. Thermal radiation affects
In hydrocarbon fuel combustion, the familiar products the structure and combustion characteristics of hydrocarbon
of combustion H2 O, CO2 , CO are particularly important and composite fuels, as well as the NO formation due to the
owing to their comparatively high absorptivities and emis- sensitivity of thermal NO kinetics to temperature. The non-
sivities in the near infrared region. The computation fluid accurate prediction of combustion phonemena in the mod-
dynamic (CFD) modelling of combustors burning hydrogen elling of hydrogen and hydrogen–hydrocarbon fuels causes
non-accurate NO predictions. For example, higher tempera-
ture predictions will cause higher NO predictions due to the
∗ Tel.: +90 352 4374901; fax: +90 352 3302740. thermal NO mechanism. Therefore, the use of a radiation
E-mail address: ilbas@erciyes.edu.tr (M. Ilbas). model for an accurate combustion simulation is essential.

0360-3199/$30.00 䉷 2004 International Association for Hydrogen Energy. Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijhydene.2004.10.009
1114 M. Ilbas / International Journal of Hydrogen Energy 30 (2005) 1113 – 1126

Nomenclature

Ak Arrhenius pre-exponential factor b oxygen reaction order


Ea activation energy, J/kmole k1 , k2 forward rate constants
p pressure, Pa k−1 , k−2 reverse rate constants
 density, kg/m3 MNO molecular weight of NO
 mixture ratio defined as air–fuel ratio i ™ ,k the molar stoichiometric coefficient for
T temperature, K species i 
Yk mass fraction of the species, k j ™ ,k the exponent on the concentration of reac-
Ri ™ ,k the reaction rate tant j 
 the rate of dissipation of turbulent kinetic Mi ™ molecular weight of species i  , kg/kmol
energy k temperature exponent
k rate constant, turbulent kinetic energy Cj ™ the molar concentration of each reactant
f correction factor for prompt-NOx model species j 
R universal gas constant

Diffusion flames have been widely applied in industrial physical constants of fluid were taken from the literature.
process systems, such as burners and furnaces. Both numer- Inlet velocities were calculated from the known mass flow
ical and experimental investigations of turbulent diffusion rates (Qfuel = 4.467 × 10−4 kg/s, Qair = 0.015536 kg/s for
(non-premixed) flames have been the subject of extensive 100% H2 , Qair = 0.01320 kg/s for 70% H2 + 30% CH4 ,
research during recent years for several gas fuels and liquid Qair = 0.007768 kg/s for 100% CH4 ) according to mixture
fuels, because they are very important for the understanding ratio ( = 1.0). For radiation calculations, absorption coef-
of the complex interactions between the turbulent flow and ficients used in the radiation model are taken 0.6 for pure
chemical reactions. Good model simulations can be used in methane, 0.50 for composite fuel and 0.45 m−1 for pure
construction of furnaces or gas turbines as well as for the hydrogen combustion, while scattering coefficients used are
prediction of pollutants such as nitrogen oxides and carbon taken 0.01 m−1 for all cases. At the combustor walls, the
monoxide emissions. heat transfer coefficient is taken 15 W/m2 K and free stream
Some detailed numerical studies on thermal radiation temperature is taken 300 K for all cases.
and radiation models may be found in literature [1–9]
for hydrocarbon fuel combustions, but the numerical 2.1. The model combustor
studies on thermal radiation and radiation models are
very limited for hydrogen combustion, and not found for The combustor modelling has been performed for three
hydrogen–hydrocarbon combustion. different fuel composition (100% H2 , 70% H2 + 30% CH4 ,
The main aim of this work is to investigate the effect of 100% CH4 ) cases. The stoichiometric mixture ratio (=1.0)
thermal radiation and radiation models on the modelling of was used for all cases. The standard k– model [12] was
hydrogen and hydrogen–hydrocarbon composite fuel com- employed for the modelling of the turbulent flow in the
bustion. In this paper, a detailed numerical experiment is combustors. The physical domain of combustors is given
carried out for two radiation models; the Discrete Transfer Fig. 1. The physical values used in combustor modelling
Radiation Model (DTRM) and the six-flux or P-1 model, are selected as: length of combustor L = 2 m, radius of
and ignoring radiation models. Experimental measurements combustor Rc = 0.3 m, radius of fuel inlet Rf = 0.005 m,
of radial temperature and NO distributions were previously radius of air inlet Ra = 0.15 m, Qfuel = 4.467 × 10−4 kg/s,
performed at the combustor exit for three different fuel com- Tfuel = 300 K, Tair = 300 K.
positions (100% CH4 , 70%H2 + 30% CH4 , and 100% H2 )
[10].
2.2. The combustion model

Reaction rate in turbulent flow is calculated from the Ar-


2. CFD Modelling of the combustor rhenius kinetic rate expression and the eddy break-up model
which are given as
Simulation of the combustor was performed by the use The Arrhenius reaction rate expression:
of a CFD package Fluent [11] which was integrated with a
  
NOx post-processor. The boundary conditions were defined j ™ ,k Ek
in terms of temperature and constant for inlets and walls. Ri ™ ,k = −i ™ ,k Mi ™ T k Ak Cj ™ exp − . (1)
RT
The gas law was used to calculate the fluid density. The other j™
M. Ilbas / International Journal of Hydrogen Energy 30 (2005) 1113 – 1126 1115

r depends on position r and direction s), local temperature,


phase function and solid angle respectively.

outlet 2.3.1. Discrete transfer radiation model (DTRM)


air The main assumption of the DTRM (the so-called the dis-
symmetry axis
fuel z crete model) is that the radiation leaving the surface element
in a certain range of solid angles can be approximated by a
Fig. 1. The sketch of the model combustor. single ray. The discrete model has some advantages; it is a
relatively simple model, the accuracy of the model can be
increased by increasing the number of rays, and it applies
The eddy break-up reaction rate (Magnussen and to a wide range of optical thickness, as well as some dis-
Hjertager model) [13] advantages; the model assumes that all surfaces are diffuse,
  m™i the effect of scattering is not included, the implementation
Ri ™ ,k = −C − , (2) assumes gray radiation, and solving a problem with a large
k ™
™i,k
i number of rays is CPU-intensive [11,14].
The equation for the change of radiant intensity, dI , along
where C is an empirical constant and C is A for reactants,
a path, ds, can be written as
and C equals to A times B for products. A and B are the
Magnussen constants (reaction mixing rates), A = 4 and dI aT 4
B=0. The eddy break-up model relates the rate of reaction to + aI = , (7)
ds 
the rate of dissipation of the reactant and product containing
eddies and is used to take the effect of turbulence on the where a, I, T and  gas absorption coefficient, inten-
reaction rate into account. The slowest (limiting) reaction sity, gas local temperature and Stefan–Boltzmann constant
rate is employed as the reaction rate for all calculations. (5.672 × 10−8 W/m2 K 4 ) respectively. If a is constant
The fluid flow and chemical reactions must be defined, along the ray, then I (s) can be estimated as
for the combustion of fuel. Oxidation of hydrocarbons can T 4
be defined by the use of the general two-step reaction mech- I (s) = (1 − exp[−as]) + I0 exp[−as], (8)

anism as [13]
  where I0 is the radiant intensity at the start of the incremen-
x+y y tal path, which is determined by the appropriate boundary
Cx Hy + O2 → xCO + H2 O, (3)
2 2 condition. The energy source in the fluid due to radiation is
then computed by summing the change in intensity along
CO + 21 O2 → CO2 . (4) the path of each ray that is traced through the fluid control
In the present work, the following global single-step re- volume.
action for hydrogen combustion is used. The radiation intensity approaching a point on a wall
surface is integrated to yield the incident radiation heat flux,
H2 + 21 O2 → H2 O. (5) qin , as

2.3. Radiation modelling qin = Iin sQn d , (9)
sQn>0
Two different radiation models; DTRM (the so-called where is the hemispherical solid angle, Iin is the intensity
discrete model) and P-1, were used for modelling of the of the incoming ray, s is the ray direction vector, n is the
radiative heat transfer in the combustor for different fuel normal pointing out of the domain. The net radiation heat
compositions. The radiative transfer equation (RTE) for an flux from the surface, qout , is then computed as a sum of
absorbing, emitting, and scattering medium at position r in the reflected portion of qin and the emissive power of the
the direction s is given as surface
dI (r, s) qout = (1 − w )qin + w Tw4 , (10)
+ (a + s )I(r, s)
ds
 where Tw is the surface temperature of the point P on the
T 4 s 4
= an2 + I(r, s )(sQs ) d  , (6) surface and w is the wall emissivity. Fluent code incorpo-
 4 0
rates the radiation heat flux, Eq. (10) in the prediction of the
where r, s, s , s, a, n, s , , I, T ,  and  po- wall surface temperature. Eq. (10) also provides the surface
sition vector, direction vector, scattering direction vec- boundary condition for the radiation intensity I0 of a ray
tor, path length, absorption coefficient, refractive in- emanating from the point P, as
dex, scattering coefficient, Stefan–Boltzmann constant qout
(5.672 × 10−8 W/m2 K 4 ), total radiation intensity (which I0 = . (11)

1116 M. Ilbas / International Journal of Hydrogen Energy 30 (2005) 1113 – 1126

2.3.2. P-1 Radiation model In methane and hydrogen–methane composite fuel com-
The P-1 model is the simplest case of the more general bustions, the formation of NO can be attributed to two chem-
P–N model, which is based on the expansion of the radiation ical kinetic process, which are thermal NOx and prompt
intensity, I , into an orthogonal series of spherical harmon- NOx , while the formation of NO in hydrogen combustions
ics. The P-1 radiation model has also some advantages and can be attributed to only the thermal NOx . Thermal NOx is
disadvantages. The P-1 model has some advantages over the formed by the oxidation of atmospheric nitrogen at high tem-
discrete model. With the P-1 model, it is easy to solve with peratures. Prompt NOx is formed by the reactions of inter-
little CPU demand, the model includes the effect of scat- mediate species at the flame front. Indeed, NO is formed via
tering. For combustion applications where the optical thick- three mechanisms in combustion systems, fuel NOx , prompt
ness is large, the P-1 model works reasonably well. The P-1 NOx and thermal NOx [14]. In hydrogen fuel combustion
model can easily be applied to complicated geometries with there is no fuel NOx , as hydrogen does not contain nitrogen,
curvilinear coordinates. The P-1 model has also some lim- the prompt NO is generated generally in the rich combus-
itations, such as, the P-1 model like the discrete model as- tion region through an intermediate chemical species HCN
sumes that all surfaces are diffuse. The implementation as- that is generated from CH in the decomposition process of
sumes gray radiation. If the optical thickness is small, there hydrocarbon fuels. These intermediate chemical species and
may be a loss of accuracy, depending on the complexity of the prompt NO are not formed in combustion of hydrogen
the geometry. The P-1 model tends to over predict radiative as hydrogen contains no carbon.
fluxes from locolazied heat sources or sinks [9,11,14]. The following well known Zeldovich [16] reaction deter-
The radiation flux, qr , for this model is obtained from the mines the thermal NO;
following equation.

1 O + N2 → N + NO, (17)
qr = − ∇G, (12)
3(a + s ) − Cs
N + O2 → O + NO, (18)
where a is the absorption coefficient mentioned above, s is
the scattering coefficient, G is the incident radiation, and C N + OH → H + NO. (19)
is the linear-anisotropic phase function coefficient (ranges
from −1 to 1). Fluent code solves the following equation
to determine the local radiation intensity in the P-1 model The mass transport equation is solved for the NO calcu-
[9,11]. lation, taking into account convection, diffusion, production
and consumption of NO and related chemical species. The
∇(
∇G) − aG + 4aT 4 = 0, (13) overall thermal NO formation rate can be calculated as [13]

where
is d[N O] 2[O](k1 k2 [O2 ][N2 ] − k−1 k−2 [NO]2 )
= . (20)
1 dt k2 [O2 ] + k−1 [NO]

= . (14)
(3(a + s ) − Cs )
The prompt NOx formation mechanism was first reported
The wall radiation heat flux is computed using the fol- by Fenimore [17]. The prompt NOx formation is signifi-
lowing equations in the P-1 model. cant in most hydrocarbon fuel combustion conditions espe-
cially, low temperature, short residence times and fuel-rich
qr Qn = −
∇GQn, (15) conditions. At present, the prompt NO contribution to the
jG total NOx formation is considerable for only pure methane
qr,w = −
. (16) combustion and hydrogen–methane composite fuel combus-
jn
tions, due to existence of carbon in the fuel. However the
Thus the flux of the incident radiation, G, at a wall is −qr,w . actual prompt NOx formation route is complex and involves
These models are detailed elsewhere [11,12,14,15]. many intermediate reactions and species, thus are costly to
compute, the prompt NOx route is generally accepted as
2.4. NOx post-processing
CH + N2 → HCN + N, (21)
In most of the combustion cases NO accounts for about
95% of the total NOx . Nitrogen oxide (NO2 ) and nitrous ox- N + O2 → NO + O, (22)
ide (N2 O) are less significant components, of the total NOx
formation. Therefore only NO formation is calculated in this
HCN + OH → CN + H2 O, (23)
work. A NOx post-processor has been used to predict NO
formation from the hydrogen combustion, methane combus-
tion, and hydrogen–methane composite fuel combustion. CN + O2 → NO + CO. (24)
M. Ilbas / International Journal of Hydrogen Energy 30 (2005) 1113 – 1126 1117

The prompt NO formation rate is calculated from the De bustion, the predicted overall temperature levels are higher
Soete [18] global model as than those for other fuel combustions. At near combustor en-
trance, z = 0.4 m, the predicted maximum temperature lev-
d[NO]pr els were about 2000 K for methane combustion, 2200 K for
dt
    70% H2 + 30% CH4 composite fuel combustion, while it
RT 2 Ea was about 2450 K for pure hydrogen combustion (see Figs.
=f ×A× [Fuel][O2 ]b [N2 ] exp . (25)
P RT 2–4. The temperature level for pure hydrogen combustion
was higher than those for methane and composite fuel com-
The source term due to thermal and prompt NOx mech- bustion. This was because of higher-energy content of hy-
anism can be calculated separately from the following ex- drogen compared with methane. The predicted radial tem-
pression in conjunction with Eqs. (20) and (25) perature profiles for three cases (no radiation, P-1 model and
d[NO] the discrete model) are very close each other at near com-
SNO = MNO . (26) bustor entrance (z =0.4 m), while significant differences are
dt
seen between the temperature profiles with and without ra-
diation models at other axial distances (z = 0.7, 1.0, and
3. Results and discussion 1.3 m). This may be because of that the combustion prod-
ucts (CO2 , H2 O, CO, etc.) are mainly produced just after the
Predictions have been done at stoichiometric mixture ratio axial distance of 0.4 m. At the combustor exit, the predicted
( = 1.0), fuel mass flow rate was of 0.0004467 kg/s for average temperature levels were about 1400 K with both ra-
all cases. Inlet air and fuel temperatures were 300 K. The diation models and about 2650 K with no radiation, while
convection wall boundary conditions have been chosen for the measured average temperature level was about 1350 K
combustor walls. (see Fig. 8c).
The predicted radial temperature profiles at different As an example, the predicted temperature contours of
axial distances (0.4, 0.7, 1.0, 1.3 m) from entrance in the pure hydrogen combustion for three cases (no radiation, P-
combustion chamber for pure methane, hydrogen–methane 1 model, discrete model) are given in Fig. 10. While the
composite fuel and pure hydrogen combustion are given predicted NO contours for three cases are given in Fig. 11.
in Figs. 2, 3 and 4, respectively. In the figures, a pre- The overall heat released is the same in all the examined
dicted high-temperature flame zone (0.0 < z < 1.0 m and cases, radiative heat transfer is responsible for reducing the
0.0 < r < 0.1 m) was located between the centre and the size of the high-temperature regions of the flame and for
combustor wall in the first half of the combustor length. shifting them towards the combustor entrance.
Temperature level is decreasing considerably from the max- The predicted radial NO profiles at different axial dis-
imum zone through the combustor exit for both radiation tances (z = 0.4, 0.7, 1.0, 1.3 m) from entrance for pure
models, while the high-temperature profiles are contin- methane, hydrogen–methane composite fuel and pure
uing for no radiation model. No significant differences hydrogen combustion are given in Figs. 5, 6 and 7, re-
are observed between the predicted temperatures with the spectively. The predicted NO profiles show similar trends
two radiation models as seen in the work of Keramida with the predicted temperature profiles, higher the tem-
et al. [2]. perature and higher the NO formation. The predicted NO
For pure methane combustion, at the axial distance emission levels with no radiation were very high for all
of 1.3 m, the predicted average temperature level was cases due to the very high predicted temperature levels
about 1100 K with both radiation models, while it was with no radiation. The predicted NO profiles at the com-
about 2500 K with no radiation model. The difference bustor exit with and without radiation are given in Fig.
between temperature profiles with and without radiation 9. At the combustor exit, the predicted maximum NO
models are getting larger towards the combustor exit. At levels without radiation were about 250 ppm for methane
the combustor exit, the average temperature levels were combustion, 1600 ppm for 70% H2 + 30% CH4 com-
about 950 K with the P-1 model, 900 K with the discrete posite fuel combustion, and 2500 ppm for pure hydrogen
transfer model and 2300K with no radiation, while the combustion. The NO results with radiation were about
measured average temperature level was about 1050 K 80 ppm for methane, 532 ppm for hydrogen–methane, and
(see Fig. 8a). 850 ppm for pure hydrogen combustions, while the mea-
For hydrogen–methane composite fuel combustion, at the surements were about 85, 500, and 700 ppm for methane,
axial distance of 1.3 m, the predicted average temperature hydrogen–methane and pure hydrogen combustions,
level was about 1200 K with both radiation models, while it respectively.
was about 1900 K with no radiation model. At the combustor Fig. 11 shows the predicted NO contours for pure hydro-
exit, the average temperature levels were about 1050 K with gen combustion with and without radiation. High NO lev-
the P-1 and the discrete transfer models and 2400 K with els were found at high-temperature regions and low NOx
no radiation, while the measured average temperature level levels at low-temperature regions. The predicted maximum
was about 1150 K (see Fig. 8b). For pure hydrogen com- NOx level with radiation was about 800 ppm, the predicted
1118 M. Ilbas / International Journal of Hydrogen Energy 30 (2005) 1113 – 1126

no radiation
2750 p-I model
2500 discrete model
2250

Temperature [K]
2000
1750
1500
1250
1000
750
500
250 z=1.3 m
0 0.05 0.1 0.15 0.2 0.25 0.3
Radial distance [m]
2750
2500
2250
Temperature [K]

2000
1750
1500
1250
1000
750
500
250 z=1 m
0 0.05 0.1 0.15 0.2 0.25 0.3
2750 Radial distance [m]
2500
2250 axial wall of the
Temperature [K]

2000 combustor
1750
1500
1250
1000
750
500 z=0.7 m
0 0.05 0.1 0.15 0.2 0.25 0.3
Radial distance [m]
2750
2500
2250
Temperature [K]

2000
1750
1500
1250
1000
750
500
250 z=0.4 m
0 0.05 0.1 0.15 0.2 0.25 0.3 z
Radial distance [m]

symmetry
axis
r
fuel jet air jet

Fig. 2. Radial temperature profiles at distances 0.4, 0.7, 1, 1.3 m from entrance (100% CH4 ).

temperature level was about 2250 K in this region. This high (from 850 to 650 ppm) towards the combustor exit with the
NOx level was because of the high temperature level in the considerable decrease of temperature levels (from 2250 to
same region. The NOx levels were decreasing considerably 1400 K).
M. Ilbas / International Journal of Hydrogen Energy 30 (2005) 1113 – 1126 1119

2750
2500 no radiation
2250 p-I model

Temperature [K]
2000 discrete model
1750
1500
1250
1000
750
500
250 z=1.3 m
0 0.05 0.1 0.15 0.2 0.25 0.3
Radial distance [m]
2750
2500
2250
Temperature [K]

2000
1750
1500
1250
1000
750
500
250 z=1 m
0 0.05 0.1 0.15 0.2 0.25 0.3
Radial distance [m]
2750
2500
axial wall of the
2250
Temperature [K]

combustor
2000
1750
1500
1250
1000
750
500
250 0 0.05 0.1 0.15 0.2 0.25 0.3 z=0.7 m
Radial distance [m]
2750
2500
2250
Temperature [K]

2000
1750
1500
1250
1000
750
500
250 z=0.4 m
0 0.05 0.1 0.15 0.2 0.25 0.3
Radial distance [m] z

symmetry
axis

r
fuel jet air jet

Fig. 3. Radial temperature profiles at distances 0.4, 0.7, 1, 1.3 m from entrance (70% H2 + 30% CH4 ).

4. Conclusions and hydrogen–hydrocarbon composite fuel combustion in a


developed combustor.
This research work numerically analysed combustion A wide range of predicted new data from the mod-
characteristics and NO emission characteristics of hydrogen elling of hydrogen and hydrogen–methane composite fuel
1120 M. Ilbas / International Journal of Hydrogen Energy 30 (2005) 1113 – 1126

2750
2500
2250 no radiation

Temperature [K]
2000 p-I model
discrete model
1750
1500
1250
1000
750
500
250 z=1.3 m
0 0.05 0.1 0.15 0.2 0.25 0.3
Radial distance [m]
2750
2500
2250
Temperature [K]

2000
1750
1500
1250
1000
750
500
250 0 0.05 0.1 0.15 0.2 0.25 0.3 z=1 m
2750 Radial distance [m]
2500
2250
Temperature [K]

2000 axial wall of the


1750 combustor
1500
1250
1000
750
500
250 0 0.05 0.1 0.15 0.2 0.25 0.3 z=0.7 m
2750 Radial distance [m]
2500
2250
Temperature [K]

2000
1750
1500
1250
1000
750
500
250 z=0.4 m
0 0.05 0.1 0.15 0.2 0.25 0.3
Radial distance [m]
z
symmetry
axis

fuel jet r
air jet

Fig. 4. Radial temperature profiles at distances 0.4, 0.7, 1, 1.3 m from entrance 100% H2 ).

combustion with no radiation, the discrete and P-1 radiation diation model is higher than those from the modelling with
models, have been obtained and presented in this paper. The discrete and P-1 radiation models. The predicted NO emis-
maximum temperature level from the modelling with no ra- sions from the modelling with no radiation are higher than
M. Ilbas / International Journal of Hydrogen Energy 30 (2005) 1113 – 1126 1121

1000 no radiation
p-I model
discrete model
750

NOx [ppm] 500

250

0 z=1.3 m
0 0.05 0.1 0.15 0.2 0.25 0.3
Radial distance [m]
4000
3500
3000
NOx [ppm]

2500
2000
1500
1000
500
0 z=1 m
0 0.05 0.1 0.15 0.2 0.25 0.3
4000 Radial distance [m]
3500 axial wall of the
3000 combustor
NOx [ppm]

2500
2000
1500
1000
500
0 z=0.7 m
0 0.05 0.1 0.15 0.2 0.25 0.3
Radial distance [m]
1000

750
NOx [ppm]

500

250

0 z=0.4 m
0 0.05 0.1 0.15 0.2 0.25 0.3
Radial distance [m] z

symmetry
axis

r
fuel jet air jet

Fig. 5. Radial NOx profiles at distances 0.4, 0.7, 1, 1.3 m from entrance (100% CH4 ).

NO emissions from both modelling results with discrete In terms of predictive accuracy, the two radiation mod-
model and P-1 model, because of higher flame temperature els have performed similarly, both showing better agree-
predictions without no radiation. ment with the experimental data at the combustor exit
1122 M. Ilbas / International Journal of Hydrogen Energy 30 (2005) 1113 – 1126

1000 no radiation
p-I model
discrete model
750

NOx [ppm]
500

250

0 z=1.3 m
0 0.05 0.1 0.15 0.2 0.25 0.3
Radial distance [m]
4000

3000
NOx [ppm]

2000

1000

0 z=1 m
0 0.05 0.1 0.15 0.2 0.25 0.3
Radial distance [m]
4000
axial wall of the
3000 combustor
NOx [ppm]

2000

1000

0 z=0.7 m
0 0.05 0.1 0.15 0.2 0.25 0.3
Radial distance [m]
1000

750
NOx [ppm]

500

250

0 z=0.4 m
0 0.05 0.1 0.15 0.2 0.25 0.3
Radial distance [m] z
symmetry
axis

fuel jet air jet r

Fig. 6. Radial NOx profiles at distance 0.4, 0.7, 1, 1.3 m from entrance (70% H2 − 30% CH4 ).

compared with no radiation cases. This indicates that ity and low computational requirements of the P-1 (the
there is no big advantage over each of radiation models so-called the six-flux) model. The CPU time consumed
in terms of predictive accuracy, except the model simplic- on a P-IV 2 GHz PC are approximately 560 min for the
M. Ilbas / International Journal of Hydrogen Energy 30 (2005) 1113 – 1126 1123

5000
no radiation
p-I model
4000
discrete model

NOx [ppm]
3000

2000

1000

0 z=1.3 m
0 0.05 0.1 0.15 0.2 0.25 0.3
Radial distance [m]
5000

4000
NOx [ppm]

3000

2000

1000

0 z=1 m
0 0.05 0.1 0.15 0.2 0.25 0.3
Radial distance [m]

5000
axial wall of the
4000 combustor
NOx [ppm]

3000

2000

1000

0 z=0.7 m
0 0.05 0.1 0.15 0.2 0.25 0.3
1000 Radial distance [m]

750
NOx [ppm]

500

250

0 z=0.4 m
0 0.05 0.1 0.15 0.2 0.25 0.3
Radial distance [m] z

symmetry
axis

fuel jet air jet r

Fig. 7. Radial NOx profiles at distances 0.4, 0.7, 1, 1.3 m from entrance (100% H2 ).

P-1 model and 720 min for the discrete transfer model overall flame temperature decreases as methane is added to
(DTRM). the fuel due to the lower-energy input and higher flame ra-
Blending hydrogen with methane causes considerable re- diation. In hydrogen fuel combustion there is no Fuel NOx ,
duction in temperature levels and thus NO emissions. The as hydrogen does not contain nitrogen, the prompt NO is
1124 M. Ilbas / International Journal of Hydrogen Energy 30 (2005) 1113 – 1126

no radiation 300
100 % CH4 flame
2750 100 % CH4 flame p-I model
discrete model
2500 experiment 250

2250
Temperature [K]

200

NOx [ppm]
2000
no radiation
1750 p-I model
150 discrete model
1500 experiment

1250 100

1000
50
750

500
0 0.05 0.1 0.15 0.2 0.25 0.3 0
0 0.05 0.1 0.15 0.2 0.25 0.3
(a) Radial distance [m] (a) Radial distance [m]
2750
70 % H2 + 30 % CH4 flame 2000
2500 70 % H2 + 30 % CH4 flame
1800
2250
Temperature [K]

2000 1600

1750 1400
NOx [ppm]
1500 1200

1250 1000
1000
800
750
600
500
0 0.05 0.1 0.15 0.2 0.25 0.3 400
(b) Radial distance [m] 200
0 0.05 0.1 0.15 0.2 0.25 0.3
2750 (b) Radial distance [m]
2500 100 % H2 flame
3000
2250 100 % H2 flame
Temperature [K]

2000 2500
1750

1500
2000
NOx [ppm]

1250
1500
1000

750 1000
500
0 0.05 0.1 0.15 0.2 0.25 0.3
500
(c) Radial distance [m]
0
Fig. 8. Radial temperature profiles at the combustor exit for dif- 0 0.05 0.1 0.15 0.2 0.25 0.3
ferent fuel compositions. (c) Radial distance [m]

Fig. 9. Radial NO profiles at the combustor exit for different fuel


compositions.
generated generally in the rich combustion region through
an intermediate chemical species HCN that is generated
from CH in the decomposition process of hydrocarbon fu-
els. These intermediate chemical species and the prompt elling has produced a better agreement between numerical
NO are not formed in combustion of hydrogen as hydrogen predictions and experimental data. The use of both radi-
contains no carbon. The important NOx mechanism is the ation models has resulted to a substantial decrease in the
thermal NOx mechanism in combustion of hydrogen. main combustion zone especially towards the combustor
The results have confirmed that the effect of thermal ra- exit.
diation is important on flame temperature predictions. The The effect of thermal radiation on methane combustion
inclusion of radiative heat transfer in the combustion mod- was higher than the effect of thermal radiation on hydrogen
M. Ilbas / International Journal of Hydrogen Energy 30 (2005) 1113 – 1126 1125

no radiation 0.3
0.3 no radiation

0.25
0.25

0.2

Radius [m]
Radius [m]

0.2

0.15
0.15

0.1
0.1

0.05
0.05

0
0 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
(a) Axial Distance [m]
(a) Axial Distance [m]

p-I model 0.3


0.3 p-I model

0.25
0.25
0.2
Radius [m]
Radius [m]

0.2
0.15
0.15
0.1
0.1 7

0.05
0.05

14 0
0 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
(b) Axial Distance [m]
(b) Axial Distance [m]
discrete model 0.3
discrete model
0.3
0.25

0.25
0.2
Radius [m]

0.2
Radius[m]

0.15
0.15
0.1
0.1
0.05
0.05
0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 (c) Axial Distance [m]
(c) Axial Distance [m]
Fig. 11. NOx contours for different radiation models (100% H2 ).
Fig. 10. Mean temperature contours for different radiation models
(100% H2 ).

combustion due to CO2 emission in methane combustion in References


addition to H2 O emission.
[1] Morvan D, Porterie B, Loraud JC, Larini M. Numerical
simulation of a methane/air radiating turbulent diffusion flame.
Acknowledgements Int J Numer Methods Heat Fluid Flow 2000;10(2):196–227.
[2] Keramida EP, Liakos HH, Founti MA, Boudouvis AG,
The financial support from Erciyes University, Turkey un- Markatos NC. Radiative heat transfer in a natural gas-fired
der Contract number FBA.03.50 is gratefully acknowledged. furnaces. Int J Heat Mass Transfer 2000;43:1801–9.
1126 M. Ilbas / International Journal of Hydrogen Energy 30 (2005) 1113 – 1126

[3] Meunier PH, Costa M, Carvalho MG. On NOx emissions [10] Ilbas M, Yilmaz I, Kaplan Y. Investigations of hydrogen and
from turbulent propane diffusion flames. Combust Flame hydrocarbon composite fuel combustion and NOx emission
1998;112:221–30. characteristics in a model combustor. Int J Hydrogen Energy
[4] Song G, Bjorge T, Holen J, Magnussen BF. Simulation of (accepted for publication 2004), in press.
fluid flow and gaseous radiation heat transfer in a natural [11] Fluent Incorporated, Fluent 6.1.22, User’s Guide, 2003.
gas-fired furnace. Int J Numer Methods Heat Fluid Flow [12] Launder BE, Spalding DB. Lectures in mathematical models
1997;7(2/3):169–80. of turbulence. London, England: Academic Press; 1972.
[5] Al-Omari SB, Kawajiri K, Yonesawa T. Soot processes in [13] Magnussen BF, Hjertager BH, On mathematical modelling
a methane-fueled furnace and their impact on radiation of turbulent combustion with special emphasis on soot
heat transfer to furnace walls. Int J Heat Mass Transfer formation and combustion. 16th symposium (International) on
2001;44:2567–81. combustion. PA, USA: The Combustion Institute Pittsburg;
[6] Dembele S, Wen JX. Investigation of a spectral formulation for 1976. p. 719.
radiative heat transfer in one-dimensional fires and combustion [14] Ilbas M. Studies of ultra low NOx burner, PhD thesis,
systems. Int J Heat Mass 2000;43:4019–30. University of Wales, Cardiff, UK; 1997.
[7] Guo H, Ju Y, Maruta K, Niioka T. Radiation extinction limit [15] Siegel R, Howell JR. Thermal radiation heat transfer.
of counterflow premixed lean methane–air flames. Combust Hemisphere Publishing Corporation; 1992.
Flame 1997;109:639–46. [16] Zeldovich YAB. Oxidation of nitrogen in combustion.
[8] Chan SH. The role of radiative transfer in combustion. Academy of Science, USSR Institute of Chemical Physics,
Proceedings of the second international symposium on 1947.
Radiation Transfer (Radiation Transfer-II), Kuşadasi, Turkey, [17] Fenimore CP. 13th symposium (international) on combustion.
1997. p. 463–74. Pittsburg: The Combustion Institute; 1971.
[9] Gosman AD, Lockwood FC. Incorporation of a flux model [18] De Soete GG. Overall reaction rates of NO and N2
for radiation into a finite difference procedure for furnace formation from fuel nitrogen. 15th symposium (international)
calculations. 14th Symposium (International) on Combustion. on combustion. PA, USA: The Combustion Institute Pittsburg;
The Combustion Institute Pittsburg; 1973. p. 661. 1974. p. 1093.

You might also like