Você está na página 1de 13

Engineering Applications of Computational Fluid Mechanics Vol. 1, No.4, pp.

337349 (2007)

SIMULATION OF A CONFINED TURBULENT


NONPREMIXED PILOTED METHANE JET FLAME
A. L. De Bortoli
UFRGS-IM/DMPA-Department of Pure and Applied Mathematics
Bento Gonalves 9500, P.O. Box 15080, Porto Alegre, Brazil
E-Mail: dbortoli@mat.ufrgs.br
ABSTRACT: The present work develops a low cost numerical method for the solution of nonpremixed piloted methane
jet flames. This method is based on the mixture fraction for fluid flow and on unsteady flamelet models, combined with
the presumed probability density function, for the chemistry. Numerical tests, for the governing equations discretized by
the finite difference and solved by the Gauss-Seidel scheme, were carried out for turbulent, nonpremixed, nonreacting
propane-jet flow and for confined Sandia C and D flames for reasonable values of gaseous hydrocarbon chemistry. The
methodology, developed for low Mach number flows and based on a density approximation, allows to decrease
considerably the computational time while obtaining results which contributes to a better understanding of the
complexity involved in the numerical solution of piloted methane jet diffusion flames.
Keywords:

Sandia flames, low Mach number, finite difference, flamelet, LES

2002). Mahesh et al. (2006) and Pierce and Moin


(2004) simulated flames of natural gas assuming it
to be pure methane in the simulations, for example.
Due to the necessity of simplifications, the
development of appropriate models for burner's
design becomes important. But, when the burners
are improved, unexpected problems can appear
(Poinsot and Selle, 2005).
Most of the applications of technical interest are
classified as nonpremixed and turbulent (Peters,
1997 & 2000; Warnatz, Maas and Dibble, 2001;
Poinsot and Veynante, 2005); some of them include
liquid fuel injection inside a chamber and to have
high burning intensity, the fuel and the oxidizer
must be well mixed. In practice, they are not
perfectly premixed before burning and therefore,
the combustion process turns less efficient.
To model nonpremixed flames it is necessary to
have a good understanding of the combustion
process and of turbulent mixing because the
reaction takes place when the fuel and the oxidizer
mix in a molecular level (Pitsch and Fedotov, 2001;
Veynante and Vervisch, 2003). Mixing is
intensified by flame-vortex interactions (Renard
et al., 2000 ) and the heat release distribution exerts
a significant influence on the flame evolution and
on turbulence, and in regions of high burning levels
the eddy life-time is short (Peters, 2000; Warnatz,
Maas and Dibble, 2001; Baurle, 2004).

1. INTRODUCTION
Combustion theory is one of the most elegant areas
of classical phenomenology, presenting a wide
range of natural phenomena that can be deduced
from a few fundamental principles (Buckmaster
et al., 2005). In combustion there is a strong
coupling among transport (heat transfer, molecular
diffusion, convection, turbulent transport) and
chemistry and hence is a multidisciplinary topic of
research. Moreover, combustion models may turn
very complex: the reaction mechanism of iso-octane
oxidation includes 3600 elementary reactions
among 860 chemical species, with 9 of 25 reaction
classes sufficient to simulate many applications at
high temperature (Curran et al., 2002). The
mechanism of methane combustion has been
identified as having more than 300 elementary
reactions and over 30 species (Liu et al., 2003). For
fuel jet A (80% n-decane and 20% 1,2,4-rrimethylbenzene), the chemical mechanism contains
approximately 1000 elementary reactions among
100 chemical species (Mahesh et al., 2006).
However, it is hard to believe that all these species
and elementary reactions are necessary to obtain a
reasonable approximation of the flow inside a
burner (Peters and Rogg, 1992). In this way,
simplified mechanisms are usually adopted to
describe the combustion process (Apte and Yang,
Received: 7 May 2007; Revised: 11 Jul. 2007; Accepted: 13 Jul. 2007
337

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 4 (2007)

Turbulent mixing plays an important role in


nonpremixed combustion. It changes the density,
the temperature, the heat capacity, the molar mass
and also the mixture transport properties
(Dimotakis, 2005). To understand the turbulent
field it is necessary to have an accurate prediction
of the turbulent velocity field.
For nonpremixed combustion a jet flame is the most
common configuration. In many of these
applications the fuel is issued as a turbulent jet, with
or without swirl (Peters, 1992). We will consider a
jet issuing from a round nozzle with diameter d and
exit velocity Uo into a surrounding air stream which
may have a constant coflow velocity U < Uo, as
shown in Fig. 1.

Fig. 1

2. MODEL FORMULATION
Usually for piloted methane jet flames, the Mach
number is low, the pressure remains almost constant
and the heat losses to the walls are small (Peters,
2000; Poinsot and Veynante, 2005). These allow
some model simplifications.
The flamelet concept, which covers a regime in
turbulent combustion where chemistry is fast, can
be employed in most practical situations, such as
the nonpremixed combustion when the flame is
almost in equilibrium. Flamelets are thin reactivediffusive layers embedded within an otherwise
nonreacting flow field (Peters, 2000).
The stationary flamelet model Yi /t = 0 , with Yi
being the mass fraction, has the advantage that
flamelet profiles can be pre-computed and stored in
a database called flamelet library containing all
the required complex chemistry. The stationary
laminar flamelet model has been applied to
engineering calculations because of its simplicity.
Such assumption indicates that the value of the
scalar dissipation rate varies slowly; it is valid in a
jet flame till x~30D, where x is the axial coordinate
and D the jet diameter.
The unsteady flamelet calculations can be
performed using a separate code to solve the system
of parabolic equations for mass fractions interacting
with a CFD code. The main advantage of the
flamelet concept is the fact that chemical time and
length scales need not be resolved in a
multidimensional CFD code (Peters, 1998).
The mixture fraction is an important quantity in the
theory of nonpremixed combustion since it is a
conserved scalar (Peters, 2000). We write the global
reaction equation for complete combustion of a
hydrocarbon fuel as

Burner sketch.

v F [F ] + v O2 [O2 ] vCO2 [CO2 ] + v H 2O [H 2 O ] + heat

For longitudinal confined jet flames, the mixing


process is influenced by the velocity ratio between
the jet and the coflow, and the diameter of the fuel
jet. For practical duct type combustors, usually one
supplies more air than required to have a
stoichiometric mixture so that the flame is short
enough and the combustion is complete (Kanury,
1975).
In the following section, some basic concepts
related to the mixture fraction, the presumed
probability density function, the scalar dissipation
rate, the heat capacity as well as the low Mach
number, incompressible and the compressible
formulations are introduced.

(1)

The reaction equation relates the fuel and


the
oxidizer
mass
fractions
by
dYO2 /(vO2 WO2 ) = dYF /(v FWF ) . For a homogeneous
system this equation may be integrated resulting in
vYF YO2 = vYF,u YO2 ,u , where v = vO2 WO2 /(vFWF )
and subscript u corresponds to an unburnt quantity.
Consider a system of two components: the subscript
1 denoting the fuel and 2 the oxidizer. The mixture
fraction is defined as the local mass fraction of all
elements within the mixture Z = m1 /(m1 + m2) . The
local mass fraction of the unburnt fuel is
338

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 4 (2007)

YF,u = YF,1 Z
YO2 ,u

and that of the oxidizer is


= YO 2, 2 Z . The mixture fraction at any state of

~~
( u~i ) ( ui u j )
1 p

+
= 2
+
x
xj
t
x j

M
i

combustion, after inserting the fuel and the oxidizer


mass
fractions
into
equation
vY F YO2 = vY F,u YO2 ,u and its corresponding

value at stoichiometric condition, vYF = YO 2 , is


given by Z =

vYF YO2 + YO2, 2


vYF,1 + YO2, 2

.
y

A probability density function pdf represents a


probability distribution in terms of integrals. The
presumed shape pdf approach seems to be the most
appropriate method to be used together with the
flamelet equations (Peters, 2000). There are many
alternatives to a general class of shapes for the
probability density functions of the conserved
scalar, but it is desirable to have a small number of
parameters (Williams, 1985).
As the pdf provides the statistical information about
the variables, it has the ability to treat finite-rate
chemistry
and
the
turbulence-chemistry
interactions. Both the mixture fraction Z and the
scalar dissipation rate , which measures the local
diffusion zone thickness, fluctuate in turbulent
flows and their statistical distribution needs to be
~
considered. If the joint pdf P (Z, st ) (where st is
at the stoichiometric condition) is known, the Favre
~
mean of mass fraction Yi can be obtained
~
from Yi =

Y (Z,
i

st )P (Z, st ; x,t)d st dZ

~ ij
ij +
R

e x j

Mixture fraction
~
~
~ ~
( Z) ( u j Z)
T Z
+
=
t
x j
x j Sc x j

(2)

(3)

Mixture fraction variance


~
~ ~ 2

( Z " 2 ) ( u j Z " )
=
+
x j
x j
t

T Z~" 2

S c x j

+ 2 DT ( Z~)2 ~

(4)
y

Enthalpy

~
~~
( h ) ( u j h )

+
=
t
x j
x j
y

T h~

Pr x j

(5)

Species and mean species mass fraction


Z e ( 1 )

Yi
a 2Yi

= DaYF YO e 1 ( 1 )
2
t
2 Le Z

(6)

~
~
Yi (x j ,t) = Yi (Z,t)P(Z; x j ,t)dZ

(7)

Mean temperature (which is obtained from the


enthalpy)

~
h=

0 0

Y~h (T)
~

i i

(8)

i =1

Inside a burner the Mach number is usually low;


therefore, one needs numerical techniques that
solve the original compressible flow equations, but
which can also be efficiently used at low Mach
numbers. The common low Mach number
formulations of variable density are that when
acoustic waves are filtered, the density change due
to temperature variations remains independent of
the pressure. Then, one can decouple the pressure
field of the state equation from the pressure
gradients of the momentum equations.
2.1

Momentum/Navier-Stokes

where u is the velocity, the density, t the time, p


the pressure, the stress tensor, Z the mixture
fracton, DT the turbulent diffusivity, h the enthalpy,
Yi the mass fractions, T the turbulent viscosity and
T the temperature. Here the density is obtained
using the relation =

p
, 0.05 < < 0.8
~
/ (1 )T + 1

(Rutland and Ferziger, 1991), which would


decrease the computational time needed to obtain
the results by an order of magnitude.
The turbulent viscosity is given by the Smagorinsky
~
~
model T = 2 (Cs )2 S . The relation for Z "2 ,

Governing equations

The set of governing equations in nondimensionalized form can be derived by applying a


spatial, density-weighted filter resulting in:

~
~2
Z "2 = C Z 2 Z , can also be used to determine

dynamically the Cs coefficient.


In these equations (Eqs. 28), Re is the Reynolds, Sc
the Schmidt, Pr the Prandtl, Da the Damkhler, Le
339

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 4 (2007)

the Lewis and Ze the Zel'dovich numbers;


T Tu
= b
~ 0.8
Tb

and

T Tu
=
Tb Tu

(Steiner

mesh resolution is used; however, refining grid is


restricted as a result of the rapid increase of the
computational cost.
As the turbulence can affect the flow in profound
ways, it is common to find significant differences
between the DNS and LES predictions. This leads
to the possibility to use LES in coarse meshes as a
tool for determining the gross features of the flow
(Ferziger and Peri, 1999).

and

Bushe, 1998). a = 2(Z )Z st (1 Z st ) appears after the


nondimensionalization of the species mass fraction
equation.
At low Mach numbers the spatial variations in
pressure are small compared with the pressure itself
and may be neglected in the equation of state,
where p may be approximated by a constant (Lin,
1991). The temperature is obtained from the
enthalpy using a simple Newton iteration and the
density comes from the state relation.
Observe that the transformation introduced to
obtain
the
flamelet
equations
(species
equations)

Z
=
+
,
=
+
x1 x1 Z
t t Z

Z
=
+
xk Z k xk Z

eliminates

the

2.2.1 Pressure calculation using the Poisson's


equation
It seems to be natural to derive a Poisson pressure
equation from the momentum (Navier-Stokes)
equations. The mass conservation equation can be
used to simplify it, resulting in Cartesian
coordinates (Ferziger and Peri, 1999; de Bortoli,
2003)

and

convective

xi

(nonlinear) terms; besides it transforms the threedimensional problem to a one-dimensional problem


(Peters, 1984 & 1986).
2.2

x
i

xi

(ui u j ij
x j

( bi ) 2
+ 2 (9)
) +

xi
t

Such equation is frequently solved by explicit


methods; the implicit methods are usually adopted
in the case of steady flows. Based on this equation,
an algorithm to solve the Navier-Stokes equations
could be written as:

Solution procedures

For the simulation of combustion flows of technical


interest RANS, Reynolds Averaged Navier-Stokes,
and LES, Large-Eddy Simulation, seem to be good
alternatives. The DNS, Direct Numerical
Simulation, on the other hand, is the most expensive
technique because the number of grid points in each
coordinate direction increases proportionally to the
Reynolds value, limiting the DNS applications.
The subgrid modeling of small-scale turbulence is
important in LES especially at high Reynolds
numbers and when relatively coarse grids are being
used in strained flow regions. Eddy-viscosity,
dynamic and hybrid models were developed for
doing that (Riesmeier, 2003; Haworth, 1999).
Among the subgrid models the Smagorinsky is still
widely used, mainly because of its simplicity. It can
be derived based on the assumption that in the SGS
kinetic energy the production is equal to dissipation,
and they have much larger magnitude than the
transport term. The numerical application of
Germano model can lead to numerical problems
owing to the possible occurrence of negative
turbulent viscosity, because the model parameter it
produces is a rapidly varying function of the spatial
coordinates and time. It has been observed that the
model for T becomes less important when better

obtain the velocities using the momentum


equations (without the pressure term)

solve the Poisson equation for the pressure

correct the velocity field

Other possibility is to employ a fractional step


method. In this case, the approximated velocities
(ui)* are advected using the pressure field from the
previous time-step. Then, half of the old pressure
gradient is removed and one calculates a new
approximation for velocities (ui)** as follows
(where means a finite difference approximation):
(ui ) * * (ui ) * 1 p n
=
t
2 xi

(10)

and in order to satisfy the mass conservation a


Poisson's equation is needed.

xi

p n +1 2 (ui )** 2
=

x t
xi
i

(11)

Then the velocities can be recalculated using the


new pressure field. This procedure is preferred to
340

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 4 (2007)

solve unsteady flows, while SIMPLE type methods


are more frequently used to solve steady flow
problems (Ferziger and Peri, 1999).
Other possibility (Pitsch and Steiner, 2000) is to
obtain the velocities (without the pressure term)
first. Then one solves the Poisson's equation for the
pressure (Boersma, 1998) as
(ui ) * *
2 p = t n
+

t
xi

scheme after an adequate implementation of the


boundary conditions.
2.2.3 Boundary conditions
Consider the longitudinal section of the burner as
shown in Fig. 2. The boundary conditions can be
summarized as follows:
y

(12)

~
~
~
~
p Yi h Z Z "2
u~i =
=
=
=
=
=
=0
n n n n n
n

and correct the velocities with the pressure gradient.


Observe that here the pressure changes are due to
mass conservation variations.
It seems that the employment of a Poisson's
equation for the pressure is permissible. The
dynamic pressure is very small (less than 0.01% of
the total pressure) and the corresponding pressure
gradients are even smaller, but they are responsible
for local effects. Reynolds (Reynolds and Kassinos,
1995) regarded the non-local effect of pressure a
greater challenge to turbulence modeling than the
non-linearity of the Navier-Stokes equations.

except at fuel injection point where u~ = 1


~

(parabolic), YF = 1, YO2 = 0, h = C PT

and the

other conditions are given by Schefer for the


nonreacting flow or by Barlow and Frank for
Sandia flames C and D (Barlow and Frank,
2003).
y

The ability to model flows through/over complex


geometries is one of the main challenges of CFD,
Computational Fluid Dynamics, since the majority
of flows involve some kind of complex geometry.
As an alternative to the boundary fitted method, the
virtual boundary technique maintains the efficiency
of the Cartesian solution procedure (Verzicco et al.,
1998; von Terzi et al., 2001) and allows transferring
the influence between each boundary point to its
neighboring points. The force field can be smoothed
in the neighborhood of the boundary grid nodes by
interpolation (Saiki and Biringen, 1996).
The finite difference approximations of second
order, for the first and second space derivatives,
considering a one-dimensional problem, are
indicated in the following:

For the outflow the convective condition results


in:

+ uc

=0

~ ~

~~ ~

~ Z Z "2 h Y T ]T and
with p = 1 , where = [ u~ v~ w
i

u c is the convective velocity.


solid wall

inflow

fuel jet

outflow

hot pilot
coflow

solid wall

For the first derivative:

Fig. 2

Boundary conditions.

For the second derivative:

2.2.4

Stiffness

yi ~ (yi +1 2 yi + yi 1)/h 2

The loss of computational efficiency can be seen as


a stiffness problem since it is a direct consequence
of flow and acoustic speeds being widely different
(Wang and Trouv, 2004). The numerical methods
which are better for solving stiff problems do more

yi ~ (yi +1 yi 1)/ 2h
y

For the inflow it results in:


p
~
~
~
~ = Y~ = Y~
v~ = w
= 0; YO2 = 1, T = 1, = 1
F
CO 2 = YH 2 O =
x

2.2.2 Finite difference approximation in


Cartesian meshes

For the solid walls the condition results in (n


corresponds to the normal direction):

where h corresponds to the grid size.


The equations approximated using the finite
difference can be integrated using the Gauss-Seidel
341

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 4 (2007)

they are employed to present some of the following


results.

work per step, but can have larger steps, such as the
implicit methods; an alternative is the Gauss-Seidel
with some relaxation.
The exponential source term, of the Arrhenius type
approximation e-Z/T, depends strongly on the
temperature, which in combustion varies
considerably. Such variation must be transferred to
the density, which affects the stability of the
numerical code. Therefore, some relaxation have to
be employed in the density evaluation. Besides, the
dynamic pressure variations are very small and yet,
important and need to be correctly evaluated.
Decreasing reaction exponents, as often done in
many reduced schemes, leads to increased stiffness.
In the reaction rate expression the exponential term
e Ta /T can be replaced by e / e (1 ) /[1 (1 )] ,
where = (T2 T1 ) / T2 ~ 0.75 and = Ta / T2 ~ 8
(Poinsot and Veynante, 2005).

3.1 Results for turbulent nonpremixed,


nonreacting, propane-jet flow
The schematic diagram of the burner longitudinal
section is shown in Fig. 3. The duct has a square
cross section with H=1 (which corresponds to
30 cm) and the jet of propane is injected from a tube
with d=0.025H; the length of the combustor is
L=11H.

fuel jet
coflow

3. NUMERICAL RESULTS

L/100

The jet flame was chosen because it seems to be a


representative of the class of nonpremixed flames.
When showing the results (where Z, fluctuate) it
is common to show the mean values, and hence

Fig. 4

Fig. 3

Geometry for nonpremixed, nonreacting


propane jet-flow.

Comparison of the mean mixture fraction and its variance (left) and the velocity (right) profiles along the duct
centerline for a turbulent, nonpremixed, nonreacting propane-jet flow with experimental data from Schefer and
with the theoretical solution for a turbulent jet diffusion flame (Peters, 2000).

342

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 4 (2007)

discrepancies when compared to the experimental


data.
Fig. 5 presents the comparison among the
experimental/numerical/theoretical mean velocity
profiles at positions x/D = 15 and 30. The
discrepancies, which were already indicated on the
axial profiles, become more evident in the radial
profiles.

Fig. 4 shows the mean velocity (right) and the


mixture fraction (left) comparisons with
experimental
data
of
Schefer
and
theoretical/analytical
values
(Peters
and
Donnerhack, 1981; Peters, 1992 & 2000) for
turbulent jet diffusion flame. It is a round turbulent
jet of propane into coflowing air. Although the
theoretical/analytical solution was not obtained for
exactly the same situation of the experiment, it
helps to understand the general solutions behavior.
The numerical/theoretical results show small

Fig. 5

Comparison of the mean radial velocity profiles for x/D = 15 and 30 for a turbulent, nonpremixed, nonreacting
propane-jet flow with experimental data from Schefer and with the theoretical solution for a turbulent jet diffusion
flame (Peters, 1992 & 2000).

Consider the longitudinal section of the burner as


shown in Fig. 6. The duct has a square cross section
with H = 1 and a cylindrical tube which injects fuel
with d = 0.025; the tube of the coflow has a
diameter of D = 0.0267 and the burner length is
L = 11. The number of grid points was taken as
41x41x149 in the (y, z , x) directions, respectively.

3.2 Results for the burner based on Sandia


flames C and D
Sandia flames C and D consist of a main jet with a
mixture of 25% of methane and 75% of air. This jet
is placed in a coflow of air and the flame is
stabilized by a pilot. The fuel is premixed with air
in order to minimize the formation of polyciclic
aromatic hydrocarbons and soot (Pitsch, Riesmeier
and Peters, 2000). The jet velocites are 29.7 and
49.6 m/s for flames C and D, respectively (Barlow
and Frank, 2003; Schneider et al., 2003). The pilot
bulk velocities are 6.8 m/s for flame C and 11.4 m/s
for flame D.
343

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 4 (2007)

D d

Fig. 7 displays the comparison between the


experimental and numerical instantaneous mixture
fraction profiles for flames C and D along the
burner centerline. Fig. 8 shows the comparison
between the experimental and numerical
instantaneous temperature profiles for flames C and
D along the burner centerline. The discrepancies are
due to the model simplification.

fuel jet
hot pilot
coflow

L/100

Fig. 6

Longitudinal section of the burner.

Fig. 7

Comparison of the mixture fraction profiles for Sandia flames C (left) and D (right) along the burner centerline
with experimental data (Barlow and Frank, 2003; Schneider et al., 2003).

amplified by some type of dissipation which can


also be influenced by the time-step employed.
Observe that products such as NO, H2, OH, CO, etc,
are not considered in the present model.
Finally, Fig. 11 compares the radial velocity
profiles for flame D at x/D = 15 and 45. The
numerical solution is dissipative at the base of the
profile for x/D = 15; for x/D = 60, the dissipation
occurs in the proximity of its tip. The longitudinal
velocity profile agrees reasonably with the
experimental data, similar to the mixture fraction
agreement shown in Fig. 7. The results seem to be
reasonable compared to other works found in the
literature (Demiraydin, 2002; Riesmeier, 2003;
Sheikhi et al., 2005).
It has been observed that for low Mach number
flows the error can contaminate the pressure

The solution indicates the axial decreasing behavior


of the mixture fraction; the temperature increases in
the reaction zone, as expected. The temperature is
overpredicted mainly in the range x/D = 4060; the
global agreement seems to be reasonable.
Fig. 9 compares the fuel CH4 mass fraction profiles
with the experimental data for Sandia flames C and
D. The fuel consumption is very well captured in
both flames; the oscillations refer to the
instantaneous values.
Fig. 10 shows the oxidizer O2 and the product CO2
mass fractions along the burner centerline.
Although these mass fractions are well predicted in
the rich part of the flame, the oxidizer mass fraction
is underpredicted in the lean part of the flame,
which leads to an overprediction of the products
mass fractions. It seems that such behavior is
344

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 4 (2007)

existence of high-frequency acoustic waves acts as


a severe restriction on the time-stepping increments
used to advance the fully compressible equations in
time (McMurtry et al., 1986).

gradient. Accurate gradients of reacting species


may require a grid spacing several orders of
magnitude finer than that necessary to resolve other
characteristic flow structures. Moreover, the

Fig. 8

Comparison of the temperature profiles for Sandia flames C (left) and D (right) along the burner centerline with
experimental data (Barlow and Frank, 2003; Schneider et al., 2003).

Fig. 9

Comparison of the mass fraction CH4 for Sandia flames C (left) and D (right) along the burner centerline with
experimental data (Barlow and Frank, 2003; Schneider et al., 2003).
345

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 4 (2007)

Fig. 10 Comparison of the O2 and CO2 mass fraction along the burner centerline for flame D with experimental data
(Barlow and Frank, 2003).

Fig. 11 Comparison of the velocity profiles at x/D=15 and 45 for flame D with experimental data (Barlow and Frank,
2003).

346

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 4 (2007)

4. CONCLUSIONS

REFERENCES

The present work developed a low cost efficient


numerical method for the solution of nonpremixed
piloted methane jet flames. The formulation was
based on the flamelet equations for the chemistry
and on the mixture fraction for the flow. The
numerical results for the nonpremixed, nonreacting,
propane jet flow as well as for the confined Sandia
flames C and D compare reasonably with the
available data found in the literature.
At low Mach numbers the dynamic pressure is very
small (less than 0.01% of the total pressure) and the
corresponding pressure gradients are even smaller
compared with the pressure itself and may be
neglected in the equation of state, where p may be
approximated as a constant (Lin, 1991). The
pressure gradient comes from the Poisson equation
(Eq. (12)), which corresponds to a pressure
correction by mass conservation. The temperature is
obtained from the enthalpy using a simple Newton
iteration and the density comes from the
~
relation = p /[ /(1 )T + 1] , which helps to
increase the time-step by an order of magnitude
when compared to the time-step necessary using the
~
relation = p / T , due to high temperature gradients.
The above are the main contributions of this work.
The developed method, based on the low Mach
number formulation, helps to obtain reasonable
results at low cost for confined jet diffusion flames.
Such results contribute to a better understanding of
the complexity involved in the numerical solution
of piloted methane jet diffusion flames.

1. Apte SV and Yang V (2002). Unsteady Flow


Evolution and Combustion Dynamics of
Homogeneous Solid Propellant in a Rocket
Motor. Combustion and Flame 131:110131.
2. Barlow R and Frank J (2003). Piloted CH4/Air
Flames C, D, E and FRelease 2.0.
www.ca.sandia.gov/TNF.
3. Baurle RA (2004). Modeling of High Speed
Reacting Flow: Established Practices and
Future Challenges. 42nd AIAA Aerospace
Sciences Meeting and Exhibit, Reno, Nevada.
AIAA 2004-0267.
4. Boersma BJ (1998). Direct Simulation of a Jet
Diffusion Flame. Annual Research Briefs,
Center for Turbulence Research, 4756.
5. Buckmaster J, Clavin P, Lin A, Matalon M,
Peters N, Siwashinsky G and Williams FA
(2005). Combustion Theory and Modeling.
Proceedings of the Combustion Institute 30:1
19.
6. Curran HJ, Gaffuri P, Pitz WJ and Westbrook
CK (2002). A Comprehensive Modeling Study
of Iso-Octane Oxidation. Combustion and
Flame 129(3):253280.
7. De Bortoli AL (2003). Mixing and Reacting
Flow Simulations inside Square Cavities.
Applied Numerical Mathematics 47(3):295
303.
8. Demiraydin L (2002). Numerical Investigation
of Turbulent Nonpremixed Methane-Air
Flames. Ph.D. Thesis, Zurich.
9. Dimotakis PE (2005). Turbulent Mixing. Annu.
Rev. Fluid Mechanics 37(1):329365.
10. Ferziger JH and Peri M (1999). Computational
Methods for Fluid Dynamics. Springer-Verlag.
11. Haworth DC (1999). Large-Eddy Simulation of
In-Cylinder Flow, Oil and Gas Science and
Technology. Rev. IFP 54:175185.
12. Kanury
AM
(1975). Introduction
to
Combustion Phenomena. Gordon and Breach
Science Publishers.
13. Lin A (1991). The Structure of Diffusion
Flames. In Fluid Dynamical Aspects of
Combustion Theory. Longman Scientific and
Technical, UK.
14. Liu Y, Lau KS, Chan CK, Guo YC and Lin WY
(2003). Structures of Scalar Transport in 2D
Transitional Jet Diffusion Flames by LES.
International Journal of Heat and Mass
Transfer 46(20):38413851.

ACKNOWLEDGEMENTS
This research started at ITV/RWTH-Aachen under
the sponsorship of CAPES - Coordenao de
Aperfeioamento de Pessoal de Nvel Superior MCT/Brasil under process 0272/06-0 and continues
under the sponsorship of CNPq - Conselho
Nacional de Desenvolvimento Cientfico e
Tecnolgico under process 304600/2006-7. The
author gratefully acknowledges the financial
support from CAPES and CNPq and the
opportunity to stay at the Institute fr Technische
Verbrennung-RWTH/Aachen, Germany, from
August/2006 to February/2007.

347

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 4 (2007)

15. Mahesh K, Constantinescu G, Apte S, Iaccarino


G, Ham F and Moin P (2006). Large-Eddy
Simulation of Reacting Turbulent Flows in
Complex Geometries. Journal of Applied
Mechanics (Transactions to ASME) 73:374
381.
16. McMurtry PA, Jou W-H, Riley JJ and Metcalfe
RW (1986). Direct Numerical Simulations of a
Reacting Mixing Layer with Chemical Heat
Release. AIAA Journal 24(6):962970.
17. Peters N (1984). Laminar Diffusion Flamelet
Models
in
Non-premixed
Turbulent
Combustion. Prog. Energy Combust. Sci.
10:319339.
18. Peters N (1986). Laminar Flamelet Concepts in
Turbulent Combustion. Twenty-first Symposium
on Combustion, The Combustion Institute,
12311250.
19. Peters N (1992). Fifteen Lectures on Laminar
and Turbulent Combustion. Ercoftag Summer
School, Aachen, Germany.
20. Peters N (1997). Four Lectures on Turbulent
Combustion. RWTH-Aachen, Germany.
21. Peters N (1998). Turbulent Combustion Models
in CFD. ECCOMAS Conference 98. John Wiley
& Sons, 2837.
22. Peters N (2000). Turbulent Combustion.
Cambridge University Press.
23. Peters N and Donnerhack S (1981). Structure
and Similarity of Nitric Oxide Production in
Turbulent Diffusion Flames. Eighteenth
Symposium on Combustion, The Combustion
Institute, 3341.
24. Peters N and Rogg B (1992). Reduced Kinetic
Mechanisms for Applications in Combustion
Systems. Ed. Peters and Rogg.
25. Pierce C and Moin P (2004). Progress-Variable
Approach for Large-Eddy Simulation of
Nonpremixed Turbulent Combustion. Journal
Fluid Mech. 504:7397.
26. Pitsch H and Fedotov S (2001). Investigation of
Scalar Dissipation Rate Fluctuations in Nonpremixed Turbulent Combustion Using a
Stochastic Approach. Combustion Theory and
Modeling 5(1):4157.
27. Pitsch H, Riesmeier E and Peters N (2000).
Unsteady Flamelet Modeling of Soot Formation
in Turbulent Diffusion Flames. Combustion Sci.
and Technology 158:389406.
28. Pitsch H and Steiner H (2000). Large-Eddy
Simulation of a Turbulent Piloted Methane/Air

29.

30.
31.

32.

33.

34.
35.

36.

37.

38.

39.

40.

348

Diffusion Flame (Sandia Flame D). Physics of


Fluids 12(10):25412553.
Poinsot T and Selle L (2005). LES and
Acoustic Analysis of Combustion Instabilities
in Gas Turbines. Plenary LectureECCOMAS
Computational
Combustion
Symposium,
Lisbone, Portugal.
Poinsot T and Veynante D (2005). Theoretical
and Numerical Combustion. 2nd Edition.
Edwards.
Renard P-H, Thvenin D, Rolon JC and Candel
S (2000). Dynamics of Flame/Vortex
Interactions. Progress in Energy and
Combustion Science 26(3):225282.
Reynolds WC and Kassinos SC (1995). Onepoint Modeling of Rapidly Deformed
Homogeneous Turbulence, Proc. Roy. Soc.
London A451, 87104.
Riesmeier E (2003). Numerische Simulation der
Verbrennungsprozesse in Gasturbinen und
Mild-Brennkammern. Ph. D. Thesis, RWTHAachen, Germany.
Rutland C and Ferziger JH (1991). Simulation
of Flame-Vortex Interactions. Combustion and
Flame 84:343360.
Sagaut P, Montreuil E and Labb O (1999).
Assessment of Self-Adaptive SGS Models for
Wall Bounded Flows. Aerospace Science and
Technology 3(6):335344.
Saiki EM and Biringen S (1996). Numerical
Simulation of a Cylinder in Uniform Flow:
Application of a Virtual Boundary Method.
Journal of Comp. Phys 123(36):450465.
Schefer RW. Data Base for a Turbulent,
Nonpremixed, Nonreacting Propane-Jet Flow.
Combustion
Research
Facility,
Sandia
National Laboratories, Livermore, CA.
www.sandia.gov/TNF.
Schneider Ch, Dreizler A, Janicka J and Hassel
EP (2003). Flow Field Measurements of Stable
and Locally Extinguishing HydrocarbonFuelled Jet Flames. Combustion and Flame
135:185190.
Sheikhi MRH, Drozda TG, Givi P, Jaberi FA
and Pope SB (2005). Large Eddy Simulation of
a Turbulent Nonpremixed Piloted Methane Jet
Flame (Sandia Flame D). Proceedings of the
Combustion Institute 30:549556.
Steiner H and Bushe WK (1998). LES of
Nonpremixed Turbulent Reacting Flows with
Conditional Source Term Estimation. CTR
Annual Research Briefs, 2334.

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 4 (2007)

41. Verzicco R, Mohd-Yusof J, Orlandi P and


Haworth D (1998). LES in Complex
Geometries Using Boundary Body Forces.
CTRProceedings of the Summer Program,
171186.
42. Veynante D and Vervisch L (2003). Turbulent
Combustion Modeling. Lecture Series, du Von
Karman Institute.
43. von Terzi D, Linnick M, Seidel J and Fasel H
(2001). Immersed Boundary Techniques for
High Order Finite Differences Method. AIAA
01-2918, 19.
44. Wang Y and Trouv A (2004). Artificial
Acoustic Stiffness Reduction with Fully
Compressible, Direct Numerical Simulation of
Combustion.
Combustion
Theory
and
Modelling 8(3):114.
45. Warnatz J, Maas U and Dibble RW (2001).
Combustion:
Physical
and
Chemical
Fundamentals, Modeling and Simulation,
Experiments, Pollutant Formation. 3rd Edition.
Springer-Verlag.
46. Williams FA (1985). Combustion Theory. 2nd
Edition. Addison-Wesley, Redwood City, CA.

349

Você também pode gostar