Numerical Heat Transfer, Part A: Applications: An International Journal of Computation and Methodology
Numerical Heat Transfer, Part A: Applications: An International Journal of Computation and Methodology
To cite this article: D. Elkaim , M. Reggio & R. Camarero (1993) CONTROL VOLUME FINITE-ELEMENT SOLUTION OF A CONFINED
TURBULENT DIFFUSION FLAME, Numerical Heat Transfer, Part A: Applications: An International Journal of Computation and
Methodology, 23:3, 259-279, DOI: 10.1080/10407789308913672
Taylor & Francis makes every effort to ensure the accuracy of all the information (the “Content”) contained
in the publications on our platform. However, Taylor & Francis, our agents, and our licensors make no
representations or warranties whatsoever as to the accuracy, completeness, or suitability for any purpose of the
Content. Any opinions and views expressed in this publication are the opinions and views of the authors, and
are not the views of or endorsed by Taylor & Francis. The accuracy of the Content should not be relied upon and
should be independently verified with primary sources of information. Taylor and Francis shall not be liable for
any losses, actions, claims, proceedings, demands, costs, expenses, damages, and other liabilities whatsoever
or howsoever caused arising directly or indirectly in connection with, in relation to or arising out of the use of
the Content.
This article may be used for research, teaching, and private study purposes. Any substantial or systematic
reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any
form to anyone is expressly forbidden. Terms & Conditions of access and use can be found at http://
www.tandfonline.com/page/terms-and-conditions
Numerical Heat Transfer, Part A, vol. 23, pp. 259-279, 1993
The control volume finite-element method in conjunction with the vom'ciry stream func-
tion fonnuhtion is used to produce a numerical solution of the confined turbulent diffu-
sion JIame. The turbulence is described by the k-c model with wall functions near solid
boundaries. The cornbuslion kinetics are defined by a one-step infhite mte chemical
readion. The validiry of the numerical method is assessed by a comparison between
present numerical results and experimental datafrom the litemturefor a hbomtory turbu-
lent diffusion W e .
INTRODUCTION
The authors acknowledge the financial support provided for this research by the CRlM (Centre de
Recherche Informatique de Montrhl), the FP (Fondation de Polytechnique), Gaz Metropolitain of Quebec and
the NSERCC (National Science and Engineering Research Council of Canada).
NOMENCLATURE
constants temperature, 'K
constants velocity components in the x and y
mole fraction directions, respectively, d s
specific heat of species i at constant average velocity components in the x
pressure a n d y directions, respectively, d s
mixture fraction local velocity components in the X and
enthalpy, Ilkg Y directions. respectively, d s
heat of reaction, Ilkg Cartesian coordinates, m
turbulent kinetic energy, m2/s2 local coordinates, m
Downloaded by [Moskow State Univ Bibliote] at 21:22 07 November 2013
MATHEMATICAL FORMULATION
The equations governing steady two-dimensional axisymmetric turbulent reacting
flow may be divided into two groups for convenience: the fluid motion equations and the
species transport equations with the related combustion model. Although radiation heat
transfer may be important in such flows (especially for large industrial furnaces), it is
not taken into account here.
The stream function $ is expressed in such a way that the continuity equation is
identically satisfied. We then have
rpu - a$
-
ar
rpv - --a$
ax
Using Eqs. (1) and (2). the governing equations for the fluid flow problem become
[6] as follows.
C O m D TURBULENT DIFFUSION FLAME 261
where p, is the turbulent viscosity evaluated according to the k-e model of turbulence [5]
as follows.
Dissipation rate of k
a (rpue) + -
-
ax
a (rpve)
ar
-
Gk or c,Gk are generation terms, as opposed to the destruction terms -pe and - cfle.
Gk is given by
This turbulence model has five constants, uk,a,, c,, c,, c,, for which the values
0.9, 1.22, 1.44, 1.92, and 0.09 are commonly used [5], respectively.
262 D. ELKAIM ET AL.
In general, the k-E model is only valid in regions where the flow is entirely
turbulent. Close to the solid walls, viscous effects become dominant, and such a model
does not lead to acceptable predictions. Later in the present work, the wall function
method is used to adequately describe the flow in the vicinity of solid walls.
Combustion Model
We consider only diffusion flames and a fast chemistry with a one-step irreversible
chemical reaction. It is assumed that the chemistry is sufficiently fast and that intermedi-
ate species do not play a significant role. In these physically controlled diffusion flames,
the mixture composition can be related to one conserved scalar quantity, namely, the
Downloaded by [Moskow State Univ Bibliote] at 21:22 07 November 2013
mixture fraction. In a two-feed system, the mixture fraction is conserved under chemical
reactions and is defined by
where
and m, and m,, denote the mass fraction of fuel and oxidizer, respectively, i is the
stoichiometric oxidant required to bum 1 kg of fuel, and the subscripts F and 0 denote
the fuel and air stream conditions at the inlet.
From the assumption of chemical equilibrium, there will be no oxidant present for
a mixture fraction richer than stoichiometric and no fuel present when the mixture
fraction is below stoichiometric. Neglecting PV work and radiation, and for adiabatic
operation under the assumption of a unit Lewis number, the enthalpy equation is similar
to the mixture fraction equation, and the enthalpy can be calculated directly from the
mixture fraction and inlet enthalpy values. Thus for all the combustion-related variables,
i.e., the enthalpy, the mass fraction of the fuel, the oxygen, and the products, it is
sufficient to solve the transport equation for the mixture fraction f. All the subsequent
unknowns can be deduced from it [7].
Because we have chosen to put more emphasis on the numerical method than on
the combustion model, at the present stage, we will not take into account the chemistry-
turbulence interaction. However, for a more rigorous modeling of a combustion system,
the mixture fraction variance should also be calculated [7].
The variable f is obtained by solving the following transport equation:
where r - -
pJaf and of 0.9.
In order to close the system given by Eqs. (1)-(12). other equations for the ther-
modynamical properties are needed. These are as follows.
CONFINED TURBULENT DIFFUSION FLAME
r main flow .
variation [8]. In this case, conditions on the wall are related to conditions at point p
through this law.
The computational procedure thus skips over the viscous sublayer, where many
grid points are necessary to adequately describe the steep gradients in this region. It is
only necessary to make sure that point p is indeed in the turbulent sublayer. Other
types of sublayers can be considered in building the law of the wall [9] with the point p
within any of these sublayers. Naturally, the velocity variation has to be changed
accordingly.
Here we consider only two sublayers (viscous and turbulent), with the first compu-
tational point p being in the turbulent sublayer. At this point the velocity Up is parallel to
the boundary and has a logarithmic variation ([8]):
Downloaded by [Moskow State Univ Bibliote] at 21:22 07 November 2013
where u* is the friction velocity and y;represents a dimensionless distance from point p
to the wall. They are defined by
parameter ( K - -
where 7, is the shear stress at the wall, K is the von Kannan constant, E is a roughness
0.4, E 9.7 [8]), and yp is the actual distance from point p to the wall.
It is the value of the dimensionless distance yithat sets the limits between the different
sublayers. For the turbulent layer, y; is approximately between 10 and 400 [9].
To obtain relations involving the values of k and E at point p, it is common to
suppose that the turbulent sublayer is in local equilibrium, so that the rate of production
of k is exactly equal to its destruction rate. Therefore, from Eqs. (6) and (8) and taking
into account only the planar two-dimensional terms gives
where y is the distance from the wall. Moreover, considering the zero velocity at the
wall and supposing that the velocity at point p is parallel to this (so that there is no
variation of the v velocity component in the normal direction to the wall), Eq. (18) at
point p becomes
CONFINED TURBULENT DIFFUSION FLAME 265
Now let us further suppose that the shear stress is constant in the sublayer, so that
replacing (aulay)lp from Eq. (21) into Eq. (19) and using the definition of p,, Eq. (9),
Downloaded by [Moskow State Univ Bibliote] at 21:22 07 November 2013
On the other hand, the combination of Eq. (18) with the logarithmic variation of the
velocity, Eq. (IS), and with kp described by the above relation, yields
u*~
4'- (23)
KYP
Equations (22) and (23) give the values of k and E at point p without solving the
transport equations, Eqs. (6) and (7). In order to obtain u*, Eqs. (15) and (17) are
combined, which yields the following nonlinear equation:
The velocity Up is known from the solution of the vorticity and stream function
transport equations, Eqs. (3) and (4). The values of kp and ep are used to calculate the
turbulent viscosity and serve as boundary conditions (Dirichlet) for the rest of the do-
main (Fig. 2). It is not necessary to calculate k and e at the walls. The viscosity there is
set equal to the laminar viscosity.
Inlet. Values of k or E are not known at the inlet and, if they are not given by
experimental data, some reasonable assumptions can be made. The kinetic energy of
turbulence is estimated according to a certain percentage of the square of the average
inlet velocity:
Outlet. We suppose that the flow extends over a sufficiently long domain so that
it is fully developed at the exit section. Thus for any variable @ the condition is
Symmetry axis. Here the radial derivative of the variables is set equal to zero:
where points W and P and the distance Y, are indicated on Fig. 3. Note that Eq. (29) is
Downloaded by [Moskow State Univ Bibliote] at 21:22 07 November 2013
valid for orthogonal and nonorthogonal grids, so that irregular boundaries can be cor-
rectly treated. It is also valid for variable density flows.
Even if a second-order boundary condition is more accurate, numerical experi-
ments show that a higher order formulation can lead the overall computational procedure
to divergence as the Reynolds number increases. It is therefore more common to use a
first-order boundary condition, as is done here.
For the stream function, we impose a constant value on the walls because they are
also streamlines. In order to find this value, the velocity profile can be integrated at a
section where it is known (inlet), and the value of $ can be fixed on another boundary
(symmetry axis or solid). This allows the integration constant to be calculated.
Turbulent flow. For turbulent flows, the calculation of the stream function re-
mains the same as before. However, the condition for the vorticity has to be changed to
reflect the law of the wall. In this case, we must find a way to impose the shear stress 7,
at the wall according to the logarithmic distribution of the velocity close to the wall. In
order to do this we must first recall that the mesh point adjacent to the wall is in the
turbulent sublayer (Fig. l), and seek one way of forcing the logarithmic variation close
to the wall. When the primitive (u, u, p) formulation is used, a well-known alternative is
to compute a fictitious slip velocity on the wall. It is deduced from the logarithmic
variation as follows: First, Eq. (15) is derived with respect t o y , which here represents a
local coordinate normal to the wall:
then, at point P and using backward differences, the derivative aU/ay is evaluated:
condition aflan -
which the values of the mixture fraction f can be deduced. On solid boundaries, the
0 is imposed, n being the direction normal to the wall.
NUMERICAL SOLUTION
To solve the set of partial differential equations previously described, the control
volume based finite-element technique presented by Baliga and Patankar [3] has been
followed. Each equation in the system can be cast in the following generic form:
CONFINED TURBULENT DIFFUSION FLAME 269
where 6 represents the scalar, which undergoes convection and is diffused through the
field, r the exchange coefficient, and S, a source or sink term. A description of the
discretization method by reference to this general transport equation follows. More
details may be found in Refs. [3, 12, 131.
the computational point P, a control volume is created by joining the centroids of all
neighboring elements through the midpoints of the corresponding sides (Fig. 4).
Following the finite-volume framework, Eq. (35) is integrated over the control
+.
volume. This procedure requires an interpolation function for Baliga and Patankar [3]
developed an approach based on the idea of using the exact solution of the one-
dimensional convection-diffusion equation as the interpolation function, namely, the ex-
ponential scheme proposed earlier by Patankar [14]. As an extension, an interpolation
function is introduced that is as close as possible to the exact solution of the two-
dimensional convectiondiffusion equation. With the origin located at the centroid of the
element, a locally flow-oriented coordinate system (X, Y) is defined (Fig. 5). For each
triangular element, the interpolation function for 6 is given by
with
where
r and p are average values of the exchange coefficient and the density that prevail over
CONFINED TURBULENT DIFFUSION FLAME 271
the element. The values of A , B, and C are uniquely determined by the values of 6
pertaining to the three nodes 1 , 2, 3.
The beneficial characteristics of this shape function have been further demon-
strated by Prakash and Patankar [15] and Hookey and Baliga [13, 161 for fluid flow or
heat transfer problems using the primitive variable formulation or the vorticity \ stream
function formulation [ll]. It must be pointed out that because there are no convection
terms in the partial differential equation for the stream function +, the interpolation
function for this variable is bilinear and given by $ -
ax + by + c. The resulting
numerical scheme is then applied to solve reacting turbulent incompressible flows.
Discretized Equations
Downloaded by [Moskow State Univ Bibliote] at 21:22 07 November 2013
The discretization of the equations is carried out by integrating Eq. (35) over the
defined control volume. Using Simpson's rule through points a-r-o or o-r-c (Fig. 5)
and applying the divergence theorem, an equation of the form
is obtained for each computational point with a,, a,,, and dp called the discretization
coefficients. Details on how these coefficients can be computed may be found in Refs.
[12] and [13]. The subscript nb refers to neighboring points, and p stands for the compu-
tational point P.
has to be written for each point of the boundary. Consider the boundary point W together
with the triangle containing the normal to the boundary at that point, and suppose a
linear variation of the property 4 in that triangle. Then in the local X, Y frame of Fig. 6,
b has the form
where N, are the base functions associated with the linear interpolation function of 6 in
the triangle containing the normal. Thus, if the normal gradient is known from the
boundary condition, then Eq. (41) is precisely of type Eq. (39). This is very useful when
computations are done in a totally implicit way, as is the case here.
Source terms. Source terms S(6) are first linearized in the following form:
Downloaded by [Moskow State Univ Bibliote] at 21:22 07 November 2013
where subscripts L and NL stand for the linear and nonlinear parts of the source term.
Then Eq. (42) is integrated over the control volume by assuming average values of S,
and S,, that prevail over the element. Thus
where Vi is the volume (an area in two-dimensions) of a neighbor element (nbe) i, and
the sum is taken over all the neighboring elements of the computational point. Assuming
that an average value mi within an element is given by
where the subscripts 1, 2, and 3 indicate local node numbers, it is possible to calculate
the contribution of the linearized source term to the coefficient a, of the computational
point P, and also to the coefficient a, of its neighboring nodes. This contribution of the i
element, denoted by as,,is given by
Node P and its neighbors will receive this contribution from all neighboring elements.
Note also that the minus sign in Eq. (45) appears because the discretization coefficients
of Eq. (39) are on the left-hand side, while the source term of Eq. (35) is on the right-
hand side.
Additional considerations have to be taken into account for the k and E source
terms in order to avoid overshoots in the solution (negative values of k or E). The
procedure consists in linearizing the negative part of the source term first and then
including it in the discretization coefficient a, on the left-hand side of Eq. (39). For a
reason that will become apparent later, even if the remaining positive part of the source
term can also be linearized and moved over to the left-hand side, it is left on the right-
hand side. The linearization of the negative part is done in the following way.
CONFINED TURBULENT DIFFUSION FLAME 273
Let S, denote the negative part of the source term, 4 being k or E. S; may be
written as
the left-hand side of Eq. (39) makes the resulting matrix more diagonally dominant, and
the iterative procedure converges. Obviously, the positive part must remain on the right-
hand side. Note that during the first iteration steps, the values of k and 6 remain nega-
tive, and after that they become admissible.
rpu - a~ -
- b
rpu - --all.ax - -,
However, for comparison purposes, it is often desirable to compute the nodal
velocities. In this case, a good approximation is obtained by computing a weighted-area
nodal velocity at point p, according to
Again, the sum is over all the elements neighboring the computational point P. The u
velocity component is computed in a similar way.
NUMERICAL RESULTS
The procedure and the method described above were applied to solve test cases for
(1) turbulent axisymmetric flow in a channel and (2) turbulent axisymmetric confined
diffusion flame. All test cases are accompanied by experimental or analytical data from
the literature, which is used to assess the validity and accuracy of the method.
'hrbulent flow in a circular pipe is a classical and yet essential test for validating
the setting of the turbulence model. We consider here a fluid that enters a circular
channel with a uniform velocity profile and a Reynolds number (based on the diameter
and the average inlet velocity) equal to 1.1 x 10' (Fig. 7). At the exit, the flow is
entirely developed, and we want to predict the radial velocity distribution at this station.
Figure 8 shows a comparison between our numerical results and Nikuradse's
quasi-analytical results [8]. The mesh consisted of 576 elements and 325 nodes uni-
fomdy distributed (13 x 25). Correspondence of the results is excellent.
This test case has been experimentally [I81 and numerically [19] studied. The
general characteristics of this flow are given in Fig. 9. Air and fuel, initially separated,
are injected into an axisymmetric burner, where they mix and react. Because of the
symmetry of the problem, the computational domain is only half the physical domain.
Comparisons are made between our numerical predictions and the experimental data of
the mixture fraction at several stations in the burner and of the combustion products near
the exit.
At all three first-comparison stations (Fig. lo), the value o f f is overpredicted,
especially in the mixing zone of the two jets. Part of this disagreement is attributed to the
difficulty in the adequate physical modeling of turbulent reacting flows. As stated ear-
lier, the turbulence-combustion interaction is not taken into account. Of course, at the
last comparison station for f, concordance is excellent because the mixing and reacting
CONFINED TURBULENT DIFFUSION FLAME
Downloaded by [Moskow State Univ Bibliote] at 21:22 07 November 2013
zone is passed. Closer to the solid walls, predictions are better as well. All these results
are also corroborated by the numerical predictions of Ref. [19].
Figure 11 shows the comparison of the molar concentrations of the combustion
products. There are slight differences between the predictions and the experimental data
for H,O and CO, because of the oversimplified one-step five-species chemical reaction.
Indeed, it can be seen that small traces of Ar, CO, and H, are measured experimentally.
Those species are not accounted for by the present combustion model. For the nonreact-
ing species (N, in Fig. l l ) , concordance is excellent, as expected.
CONCLUSIONS
A numerical procedure has been developed for solving turbulent reacting flow.
The velocity-pressure coupling was addressed via the vorticity stream function formula-
tion. Turbulence has been modeled using the k-e model together with wall functions.
The control volume based finite-element method has been used as the numerical method
in conjunction with the exponential interpolation function. Boundary conditions for the
vorticity have been given careful consideration when used with wall functions. To avoid
undershoots, special treatments for the k and E source terms had to be done. Physical
modeling has to be improved to enhance the overall accuracy of the resulting numerical
Fig. 8 Radial velocity distribution in a circular channel. Solid line, present predictions;
squares, Ref. [a].
D. ELKAIM ET AL.
Downloaded by [Moskow State Univ Bibliote] at 21:22 07 November 2013
Fig. 9 Global features of the turbulent confined diffusion flame and location of the comparison
stations (drawing scaled in the x direction).
REFERENCES
-
Fig. 11 Radial distribution of the species. Solid line, present predictions; squares and
circles. experimental data [18]; axial distance x 1.375 m.
Finite-Element Methods: Pan 2, Incompressible Fluid Flow Problems, Nwner. Heat Transfer,
vol. 14, pp. 273-293, 1988.
17. IBM. Engineering and Scienrific Subroutine Library, 4th ed., 1983.
18. D. L. Smoot and H. M. Lewis, 'hrbulent Gaswus Combustion: Pan I, Local Species Con-
centration Measurements, Combust. Flame, vol. 42, pp. 183-196, 1981.
19. D. L. Smoot and P. J. Smith, 'hrbulent Gaseous Combustion: Part LI, Theory and Evaluation
of Local Properties, Combust. Flame, vol. 42, pp. 277-285, 1981.