Вы находитесь на странице: 1из 12

A thermodynamic approach towards

turbulent diffusive reactive flows


Dragos Isvoranu, Mircea Marinescu, Dorin Stanciu
Dept. of Thermodynamics
University Politehnica of Bucharest
313, Spl. Independentei, sect.6
77206 - Bucharest, ROMANIA
Tel: 40-1-401.04.00/339
E-mail: dragos@mdd.combustion.pub.ro

Abstract. The paper deals with entropy generation rate modeling in turbulent diffusive
reacting flows. In the general framework of k-epsilon turbulence model we have considered
the averaged entropy transport equation from which we have extracted the mean source
term of entropy. We present for the first time the possibility of modeling the viscous,
thermal, diffusion and reactive dissipation terms, based on probability density function, for
a diffusion methane-air turbulent flame in an axial-symmetric combustor. Plane
distributions of dissipation are available as well as a comparison between global entropy
generation rate calculated on the averaged Second Law basis and by integrating local
entropy generation rate. The analysis could serve to improve combustor geometry, flow
conditions and to optimize k-epsilon model constants.

Nomenclature
~ ~
Ss mean entropy local source term fξ DF function
~
Dµ mean turbulent viscous dissipation M molar mass
~
DT mean turbulent thermal dissipation υ stoichiometric coefficient
~
DY mean turbulent diffusion dissipation cp mass specific heat
~
Dch mean turbulent chemical dissipation Hi lower calorific value
µi chemical potential of i species ζ formal random variable
χ~ mean scalar dissipation D diffusion coefficient
χ~st mean stoichiometric scalar dissipation Π viscous stress tensor
s instantaneous local specific entropy

1. Introduction
The basic theory of the entropy generation rate in turbulent reactive flows are to be found in
Isvoranu, [7], so the only goal of this paper is to show the practical way of its application.
In order to accomplish this purpose we have thought to comply with a real flame simulation
for which the experimental data was taken from Peeters [1]. The basic equations are given
under the general transport formulation:
∂ ρ v~ j φ ∂ ∂φ ~
− Γeff = Sφ (1)
∂x j ∂x j ∂x j
where diffusion coefficient and source term for every physical quantity transported are
revealed in table 1.

Table 1

Physical Sym Γeff ~



quantity bol
mass ρ 0 0
momen- v~i µ + µT (
− ∂ p ∂xi + ρ gi + ∂ Γeff ∂v~j ∂xi ∂x j )
( )
tum
- (2 3) Γeff ∂v~j ∂x j + ρ k
~

( )( )
turbulent ~ µ + µT σ k Pk − ρ ε~ − µ T σ ρ ∇ p ⋅ ∇ ρ ρ
2
k
kinetic
energy
dissipa-
tion
ε~ µ + µT σ ε (ε~ k~)(Cε P − Cε
1 k 2 (
ρ ε~ − Cε 3 (µT σ ρ ) ∇ p ⋅ ∇ ρ ) ρ
2
)
rate
~ µ PrL + µ T PrT
stagna- h Φ − ρ ε~ + v~ j ⋅ ∂ p ∂x j
tion
enthalpy
~
mixture ξ ρ D + µT σ f 0
fraction

The constants appearing in the k-epsilon equations as well as those in the mixture fraction
and mixture fraction variance equations are taken to have the standard values (Williams,
[2]). Turbulent viscosity µT , the molecular energy dissipation , Φ and the viscous stress
tensor, t ij are given by:
~
∂v~i ∂v~ j
k2
µ t = C µ ρ ~ ; Φ = t ij
ε
∂v~i
∂x j
; t ij = µ L +
∂x j ∂x i
2
− µ L ∇ ⋅ v δ ij
3
( ) (2)

For jet flows, as the one we will try to simulate, the standard k-epsilon model does not
produce the right spreading rates of momentum and scalar quantities. Sanders [3] suggested
to set C µ to 0.06 instead of 0.09. In the following we will use a somewhat more complex
model, based on the work of Launder et all. [4]. The constants C µ and C ε 2 are replaced
by:
C µ = 0.09 − 0.04 F jet C ε 2 = 1.92 − 0.0667 F jet (3)
The function F jet is zero in a Cartesian coordinate system and depends on the axial velocity
gradient and the jet width in cylindrical coordinates.
0.2
d dv~ax dv~ax
F jet = −
1
2
(4)
1.6∆v dx dx

is the jet diameter defined by the location where u~ = (u~ax − u~∞ ) , u


1 ~ is the
Here d 1
ax
2
2
~ ~
mean centerline velocity and ∆u = u ax − u ∞ is the difference between the centerline velocity
and the ambient velocity. Constants have been adopted from Abu-Ellail and Salem [5].

2. Burner geometry and boundary conditions


The Delft piloted diffusion flame burner was design to yield a stable axial-symmetric
turbulent non-premixed flame, burning methane in a co-flowing air stream. The fuel jet exit
has an inner diameter of 6 mm and the fuel pipe is about 1 m long, sufficient to establish a
fully developed turbulent flow at the nozzle. Two air streams are involved. The primary air
is issued from an annulus around the fuel nozzle with an inner diameter of 15 mm and an
outer diameter of 45 mm. The bulk velocity of the primary air is of 4 m/s while for the
methane /nitrogen mixture we have 20 m/s. Figure 1 shows a cross section through the
burner. The annular air is surrounded by a low speed co-flowing air stream of about 0.4
m./s, just sufficient to avoid external recirculation zone. This secondary air is issued by a
throat, yielding low turbulence levels at the entrance of the combustion chamber and a
reasonably flat velocity profile. The secondary air flow is kept at room temperature (295
K). The burner chamber depicted in figure 1 has 58 cm in diameter and 150 cm in length to
allow major simplifications from boundary conditions and geometrical point of view, that is
axial-symmetric flow and homogenous Neumann outlet conditions.

Figure 1
The following boundary conditions will be applied. On the burner rim and chamber wall,
~
Neumann conditions for stagnation enthalpy and mixture fraction ξ and wall Dirichlet
conditions for its fluctuation, ξ ′′ 2 , turbulent kinetic energy, dissipation rate and velocities.
For temperature, mixture fraction and its fluctuations, Dirichlet: conditions have been
applied at inlet boundary:
~ ~ ~
Tin1 = Tin 2 = Tin 3 = 295 K ξ 1 = 1 ; ξ 2 = ξ 3 = 0 ξ ′′ 2 1 = ξ ′′ 2 2 = ξ ′′ 2 3 = 0 (5)
In order to obtain a greater accuracy in numerical simulation, primary air inlet velocities, as
well as turbulent kinetic energy profiles have been measured for non-reacting jet, while
inlet dissipation rate has been evaluated from numerical simulation at the pipe outlet. In the
case of central fuel jet as well as in secondary air flow, constant velocities, turbulent kinetic
energy, dissipation rate profiles have been assumed, taking into account either the small jet
diameter or the negligible influence of walls. Inlet mixture fractions are:
Y F , 2 = 0.7686881 ; Yin , 2 = 0.2313119
(6)
YO ,1 = 0.2314 ; Yin ,1 = 0.7686
On symmetry axis, zero radial velocity and Neumann conditions are applied for all scalar
variables as well as on outlet boundary for all variables. In the former relations subscript 1
refers to the primary air flow, subscript 2 to the fuel mixture, 3 to the secondary air flow,
F stands for methane, O for oxigen and in for nitrogen.

3. Chemical model
The simplest way to model the chemistry in diffusion flames is to assume an infinitely
fast irreversible global reaction, yielding the so-called Burke-Schumann approximation
(Burke and Schumann, [6]) or flame sheet model. Of course, more complicated chemistry
can be imagined but for our purpose, that is to provide an example, the flame sheet model
offers the indisputable advantage of analytic solutions for all derivatives included in the
entropy generation rate. Hence, following Williams [2], we can express the mass fractions
of reactants and products only function of the mixture fraction.
ξ − ξ st
Y F (ξ ) = Y F , 2 H (ξ − ξ st ) (7)
1 − ξ st
ξ st − ξ
YO (ξ ) = YO ,1 (1 − H (ξ − ξ st )) (8)
ξ st
Yin (ξ ) = Yin ,1 + (Yin , 2 − Yin ,1 )ξ (9)
Y pr (ξ ) = 1 − Y F (ξ ) − YO (ξ ) − Yin (ξ ) (10)

where ξ st is the stoichiometric value of the mixture fraction, that reads:


YO ,1
ξ st = (11)
υO M O
Y F , 2 + YO ,1
υF M F
and subscript pr stands for product species. H (x ) is the Heaviside step function.
Considering for this practical purpose that total enthalpy is also a conserved scalar, then
temperature can be expressed only in terms of mixture fraction like:

T f (ξ ) + YO ,1 (1 − ξ ) for ξ ≥ ξ st
Hi
M Oυ O c p (ξ )
T (ξ ) = (12)
T f (ξ ) +
Hi
Y F , 2ξ for ξ ≤ ξ st
M F υ F c p (ξ )
where:
c p ,O c p,F c p ,O
T f (ξ ) = T1 + T2 − T1 ξ (13)
cp cp cp
represents the frozen mixture temperature.

4. Turbulent entropy generation rate


Entropy, like any other state parameter, will obey the same transport equation like (1).
As previously shown by Isvoranu [7], the local entropy source term, in the case of turbulent
diffusion flame with a univariate probability density function (PDF) approach, looks like:
~ ~ ~ ~ ~
S s = D µ + DT + D y + Dch (14)
~ 1 ~ ~ ρε~
D µ = ~ Π : ∇v + ~ (15)
T T
2 2
ρ
1

f (ξ )dξ
~ dT 1
DT = χ~ ρc p (16)
2 0 dξ T (ξ )
2
ρ
1 N

f (ξ )dξ
~ dYi Ri
D y = χ~ (17)
2 0 i =1 dξ Yi

ρ
1 N
d 2 Yi
µ i (ξ ) f (ξ )dξ
~ 1
Dch = χ~ (18)
2 0 T (ξ ) i =1 dξ 2

χ~ =
ρ
2
( )
ρD ∇ξ ′′
2 ~
= 2 D∇ ξ () 2
(19)

Based on the Burke-Schumann chemical model we can now describe all mass fraction and
temperature derivatives as follows:
YF ,2 0 ξ ≥ ξ st
dY F ξ ≥ ξ st dYO
= 1 − ξ st = YO ,1 (20)
dξ dξ − ξ < ξ st
0 ξ < ξ st ξ st
Hi
− Tf − YO ,1 ξ ≥ ξ st
dYin dT M Oυ O c p
= Yin , 2 − Yin ,1 = (21)
dξ dξ Hi
− Tf + YF ,2 ξ < ξ st
M Fυ F c p
Following Williams [2] the second derivatives of the mass fractions appearing in relation
(18) can be expressed, for a single step irreversible reaction, function of stoichiometric
coefficients and, for example, fuel second derivative:
d 2Y F YF ,2
= δ (ξ − ξ st ) (22)
dξ 2
(1 − ξ st )
In this last equation δ means the Dirac function. So, the chemical entropy source term
reads:
ρ 1 YF ,2 N

µ iυ i f (ξ st )
~
D ch = st χ~st (23)
2 Tst 1 − ξ st i =1
Regarding the mixture fraction probability density function we have chosen, for the sake of
lower computational time, the clipped Gaussian distribution:
1 ζ −µ
2

f ξ (ζ ) = α 0δ (ζ ) + α 1δ (ζ − 1) + {H (ζ ) − H (ζ − 1)}
~ 1
exp − (24)
σ 2π 2 σ
which can provide analytical integral expressions for (16), (17), (18), (20), (21), (22), (23).
Local parameters σ , µ , α 0 and α 1 are determined function of mixture fraction and
mixture fraction variance as presented by Elgobashi [8] and Lockwood [9] .

5. Simulation results
The discretization of transport equations was carried out in a finite volume approach on
a non-staggered, 50x50 nodes, non-uniform grid, following a semi-implicit numerical
scheme. More details can be found in the work of Ferziger and Peric [10]. Figures 2, 3, 4, 5
present the comparison between numerical simulation and experimental data on main
physical quantities as velocity, temperature and turbulent kinetic energy.

Figure 2
The mean flow field is quite well simulated as far as the basic k-epsilon turbulence model
can do. There are some discrepancies in the radial profiles of the axial velocity (figure 2),

Figure 3

mainly at x = .0.05 m, where the jet mixing is over-predicted, whereas the experiments
suggest an axial increase of the mean velocity. Further downstream, the computed profiles

Figure 4
are slightly narrower than the experimental ones, but the overall performance of the
turbulence model combined with Launder's round jet correction is good.

Figure 5

The centerline profile of the mean axial velocity depicted in figure 4 cannot reproduce the
experimental acceleration close to the burner rim. This is due to the fact that the burner rim
was too narrow to ensure a burner stabilized flame, so the experiments had to be carried out
with the help of a pilot flame which acts like a quickly expanding, highly viscous, low-
turbulent shield around the fuel jet. On the other hand, the turbulence model itself predicts
an almost isotropic distribution of normal Reynolds stresses which presumes uniform
turbulent mixing. Obviously, close to the burner rim, radial velocity fluctuations
responsible for turbulent transport in radial direction, are damped by the axial acceleration,
feature that numerical simulation fails to predict. The centerline turbulent kinetic energy
profile depicted in figure 5 shows the same trend like the experimental data, although the
maximum predicted is too small. Figure 3 illustrates the good correspondence between
computed and experimental temperature radial profiles, although the inspection of the
radial locations of maximum temperatures reveals a somewhat under-prediction of the
reaction zone spreading. Anyway, an experimental accuracy of ± 70 K has to be taken into
account in the overall evaluation. The overall conclusion is that the simulation code can
provide consistent data concerning the mean vector and scalar physical quantities. The next
figures illustrate the main outcome of the numerical analysis, that is computing the
turbulent entropy production rate (figure 10) with all its viscous (figure 6), thermal (figure
7), diffusion (figure 8) and chemical (figure 9) components. Now, having determined the
turbulent entropy source term, we can proceed to verify both the accuracy of our
computation and veridicality of our modeling. In order set up this goal, we shell rely on the
entropy turbulent transport equation for steady adiabatic systems (Isvoranu, [7]):
~ ~
ρ~s v ndA = S s dV (25)
Ξ Ω
(
~ ~
Figure 6 : Contours of Dµ S s . ) ~ ~
(
Figure 7 : Contours of DT S s . )

(~ ~
)
Figure 8: Contours of DY S s . (
~ ~
)
Figure 9: Contours of D ch S s .
(
~
Figure 10: Contours of log S s . )
where Ξ represents the surface characterized by normal vector n bounding the volume Ω .
As one can easily see, we restrict ourselves only to a global verification due to the complex
task of modeling the local terms containing correlation between temperature fluctuation and
their gradient or entropy fluctuation and gradient of mass fraction fluctuation arising in the
full form of equation (25). After integration, the previous relation reads:
~
Σ = S dV = ~
s
s dm − ~ s d m = S − S = ∆S
out in
(26)
Ω Ξ out Ξ in

where, Σ represents the global entropy generation rate, that is the right term in (25), S is
the entropy flux through Ξ in , Ξ out , the inlet-outlet surfaces of the system. Of course, the
mean specific entropy is computed also with the aid of the same probability density
function (24) taking into account its dependence mainly on mixture fraction. Computing the
global entropy generation rate from left hand side of (26) based on equations (14), (15),
(16), (17) and (18), and the global entropy flux from the right hand side of (26) in the
manner described above, we have obtained the following results presented in table 2.

Table 2

Σ (eq. 26; right side) Σ (eq. 26; left side)


46.02 [W/K] 38.60 [W/K]

Considering the right hand side result as a reference one, the evaluation of global entropy
production rate by local integration is under-predicted with a relative error of about 16%.
Of course, both results are affected by the chosen turbulence model, in our case the
standard k-epsilon model.

6. Conclusions
There are several important outcomes of this thermodynamic approach of reacting
flows. Besides the good correlation between experimental data and the numerical
simulation of this diffusion axial-symmetric flame, in spite of the simplicity of the chemical
and turbulence model used, the most important achievement is setting up new basis on fluid
flow analysis. Until now turbulent flows were characterized only by the scalar dissipation,
usually connected a conservative scalar namely, mixture fraction. We were able to set forth
that the concept of dissipation should be commonly related to the entropy production source
term, so that this notion gets a theoretical foundation on the Second Law of
Thermodynamics. In this way, scalar dissipation does not remain anymore a mathematical
artifice but it is strongly related to a physical meaning.
On the other hand, any thermodynamic process is defined by its efficiency, closely
related to different types of losses, that is entropy generation, as stated by the Second Law.
It becomes obviously that one could wish these dissipations were the smallest possible, but
in order to do this it is imperative to know what generates them and which is their structure
and magnitude. For example, figure 6 shows that turbulent viscous dissipation is higher
near the burner rim, where even a small recirculation zone appears, on the outer edge of the
fluid boundary between the expanding fuel jet and the main air stream and at the wall, near
outlet, which means the existence of a small inverse flow. Thermal dissipation (figure 7) is
concentrated at the inner edge of the fluid boundary between the expanding fuel jet and the
primary air stream in the flame front, while the diffusion dissipation's higher magnitude
(figure 8) lies close to the wall, near outlet border. Obviously, the chemical dissipation
(figure 9) is concentrated in a very thin region bordering the flame front, as predicted by the
chemical model, but its highest magnitude is found at flame tip at highest temperature. The
entropy generation distribution (figure 10) shows clearly that its highest values lie at the
interface between the fuel and primary air jets. Knowledge of this particular dissipation
structure can represent a powerful tool in order to alter some geometrical, inlet or boundary
conditions so that the global entropy production rate, in other words the losses, diminishes.
Practically, this means to optimize the thermodynamic process which in our case, for
example, would be reflected in obtaining the same maximum temperature but with a
smaller entropy generation.
Another possible approach set forth by the entropy generation rate analysis is
turbulence modeling optimization. In this case, we will take advantage of the positivity of
the entropy production rate term and of the Glansdorf-Prigogine theorem which calls for
minimum entropy generation rate for steady states.
Based on this assumption, Isvoranu [7] showed that a better agreement between the
right and left hand terms in equation (26) is found slightly altering the standard k-epsilon
turbulence model coefficients by entropy production rate minimization.
The final remark deals with the necessity that any numerical discretization scheme or
chemical mechanism should be investigated on its consistency and robustness based on this
Second Law approach, so that they do not produce locally negative source of entropy.
Acknowledgements
The authors, and especially Dragos Isvoranu, would like to acknowledge the cooperative
support provided by members of Thermofluid Department at TU Delft in his theoretical
investigation both during the two months stay in Netherlands and after. Special thanks are
due to prof. K. Hanjalic, dr.ir. Theo van der Meer and dr.ir. Tim Peeters.

References
1. Peeters, T.W.J. Numerical modeling of turbulent natural-gas diffusion flames. Ph.D.
thesis, TU Delft, 1955.
2. Williams, F.A. Combustion Theory. Benjamin / Cummings Publishing Co., Menlo Park,
CA, second edition, 1985.
3. Sanders, J.P.H. Scalar transport and flamelet modeling in turbulent jet diffusion flames.
PhD thesis, Technische Universiteit Eindhoven, 1994.
4. Launder, B. E., Morse, A., Rodi, W., Spalding, D.B. The prediction of free shear flows –
A comparison of six turbulence models. Technical report, NASA SP – 311, 1972.
5. Abou-Ellail, M. M., Salem, H. A skewed PDF combustion model for jet diffusion
flames. Journal of Heat Transfer, 112, pp. 1002-1007, 1990.
6. Burke, S.P., Schumann, T.E.W. Diffusion flames. Industrial and Engineering Chemistry,
20, pp. 998-1004, 1928.
7. Isvoranu, D. Contributions to chemical reacting flows optimization. Pd.D. Thesis, Univ.
Politehnica of Bucharest, 1999.
8. Elghobashi, S. Studies in the prediction of turbulent diffusion flames. Studies in
Convection, vol. 2, B.E. Launder, Ed., Academic Press, 1977.
9. Lockwood, F., C. The Modeling of Turbulent Premixed and Diffusion Combustion in the
Computation of Engineering Flows. Combustion & Flame, 29, 111-122, 1977.
10. Ferziger, J. H., Peric, M. Computational Methods for Fluid Dynamics. Ed. Springer,
1995.

Вам также может понравиться