COMBUSTION A N D F L A M E 81:325-340 (1990)
325
Counterflow Spray Combustion Modeling
G. C O N T I N I L L O * and W. A. S I R I G N A N O
Department of MechanicalEngineering, University of California, Irvine, CA 92717
A fuel spray planar counterfiow flame has been modeled by means of a low-Mach-number boundary layer approach. The model considers density variations and full coupling between the two phases by means of a hybrid
Eulerian-Lagrangian formulation, with droplet drag and vaporization. A spherically symmetric, unsteady droplet
model with variable properties has been employed. Conditions for similarity in such a two-phase, particle-laden
reacting boundary layer have been determined. Similarity solutions have been derived that are valid at any distance
from the stagnation point and for any liquid fuel concentration and droplet number density within the limits of
negligible particle-particle interactions. Results for the specific calculation of a monodisperse n-octane spray flame
are reported and discussed.
NOMENCLATURE
B
bi
BM
Br
C1,C2
CD
Cp
Di
Ea
f
fo
g
K
L
Le
(m)i
preexponential factor
stoichiometric coefficient for species
i
Spalding transfer number for mass
Spalding transfer number for energy
constants in the Clausius-Clapeyron
relationship
drag coefficient
specific heat
mass diffusivity for species i
activation energy
similarity variable
contribution to the stream function
due to mass vaporization
vector acceleration of gravity
strain rate of the outer flow
latent heat of vaporization
Lewis number
molecular weight, species i
* On leave from Istituto Ricerche sulla Combustione CNRNapoli, Italy.
Copyright ~) 1990 by The Combustion Institute
Published by Elsevier Science Publishing Co., Inc.
655 Avenue of the Americas, New York, NY 10010
m
n
P
Pr
Q
¢
R
r
Re
(R
s
Sc
Se, Sin, Sv
t
T
u
13
V
X
x
X
mass vaporization rate
droplet number density
pressure
Prandtl number
heat of combustion
energy flux into droplet interior
droplet radius
radial coordinate, droplet interior
Reynolds number
universal gas constant
transformed x component of droplet
vector position
Schmidt number
source terms in the gas equations due
to droplets
time coordinate
temperature
x velocity component
y velocity component
vector velocity
source term due to chemical reaction
x space coordinate
droplet vector position
species mass fraction
0010-2180/90/$3.50
326
Y
Y
G. CONTINILLO and W. A. SIRIGNANO
y space coordinate
species mass fraction
Greek Symbols
,y
~7
0
P
/)
exponents in the Arrhenius expression for reaction rate
ratio between specific heats
Kroneker delta
similarity independent variable
nondimensional temperature
heat conductivity
viscosity
kinematic viscosity
transformed
radial
coordinate,
droplet interior
density
transformed time coordinate
Subscripts
CX~
+
boil
crit
f
F
i
/
O
O
P
S
X
Y
infinity
positive y side
negative y side
boiling point
critical point
film
fuel
species
liquid
oxygen
initial or outer flow
products
surface
x component
y component
Superscripts
t
derivative along 7, gas equations;
along z, droplet equations
reference quantity
dimensional variable
INTRODUCTION
The counterflow flame has been given much attention in the past decade both from an experimental
and from a theoretical point of view. More than
one reason accounts for such interest. The configuration of two impinging streams can encompass
a complete spectrum of boundary conditions and
relevant flame characters that, for a gaseous system, vary from the diffusion flame (fuel from one
side, oxidant from the other) to partially premixed
diffusion flames, to the symmetric twin premixed
flames, to the simple premixed flame (fuel and
oxidant from one side, inert and/or products from
the other).
All of the mentioned configurations appear to be
one-dimensional in essence, at least in the neighborhood of the stagnation plane; this reflects the
existence of similarity solutions to the boundary
layer equations governing the problem, given the
proper boundary conditions, and with a minimum
amount of hypotheses valid in almost all of the
cases of interest [1].
The flames are generally located in a defined
region around the stagnation point, over a wide
range for the values of the parameters. Without
heat losses to an external body, such as a flame
holder, the flame stability is due to a complex
balance between convection, heat and mass diffusion and chemical reaction. For this last reason,
the diffusion counterflow flame is often referred
to as the "pure diffusion flame," in that it has,
for sufficiently high DamkShler number, no cold
zone in which partial mixing of the reactants takes
place, and is therefore best suited for the analysis of certain fundamental processes, like diffusion
flame extinction, which cannot be studied in the
classical Burke-Schumann configuration, as noted
by Tsuji [2]. The same reference provides an extensive review of the research work done for the
counterflow gaseous diffusion flame configuration.
Another reason to look into counterflowing
reacting systems is illustrated by the work of
Libby and Williams [3], in which strained laminar flamelets are studied as elements of a premixed turbulent flame. Experimental and theoretical work has been done for diffusion [1, 4, 5], partially premixed [6] and premixed [7-11] gaseous
counterflow flames, flame extinction [6], often by
means of large activation energy asymptotics, and
inhibition [12] have been addressed.
Less attention has been devoted to two-phase
systems. Graves and Wendt [13] first used an op-
COUNTERFLOW SPRAY COMBUSTION MODELING
posed jet diffusion flame in an experimental study
of laminar, pulverized coal flames, and by experimental evidence and theoretical arguments suggested that the system be still similar in essence
even when particles do not follow gas streamlines.
More recently, Puri and Libby [14] studied the behavior of single liquid fuel droplets in a counterflowing methane flame, comparing observed trajectories with predictions obtained by using a number of different models for droplet drag in an assumed flow field.
The same reasons that make the counterflow
configuration interesting in the study of homogeneous (gas) flames are valid as well for twophase (spray) flames. In particular, the counterflow spray configuration is seen here as a very
interesting model problem, in which droplet vaporization, drag, and combustion can be followed
in a well-defined region. The interaction of convection, diffusion, and chemical reaction phenomena determined the flame structure, and their relative influence can be studied by changing a few
main parameters, especially the rate of strain of
the flow. High-temperature differences and sharp
gradients are present as in any practical flame,
and boundary effects and heat losses are limited
to a minimum. In order to study fully the system
without having to confine attention to particular
regimes like flame extinction or small liquid-tovapor fuel ratios, full coupling between the two
phases and finite-rate chemistry are included in
the model, as illustrated in the following section.
Under certain conditions, not very restrictive and
identified in the foregoing, this will not affect the
similarity of the solution; it is clear, however, that
a numerical approach, successively described, is
required. Results are reported and discussed, and
areas of interest are identified for future work.
327
and droplets, in any combination, can be fed from
either side of the stagnation plane.
In this work, the steady state is considered for
the system, so that the effects of the inherent unsteadiness due to the discrete spray distribution is
neglected. The analysis is confined to a singlecomponent, monodisperse spray; the extension to
multicomponent, polydisperse sprays does not affect the following analysis other than to increase
the number of equations for the species in both the
gas-phase and the liquid phase, and the number of
characteristics for the droplet motion.
A hybrid Eulerian-Lagrangian formulation is
adopted. The problem will be reduced to one space
dimension via a similarity analysis. One droplet
will represent all those lying on the same plane,
perpendicular to the space coordinate.
Among the assumptions made are constant specific heats, equal binary diffusion coefficients, and
ideal gas mixture. More assumptions are mentioned where invoked. We use dimensional variables in the Eqs. 1-32; in the remaining equations,
the same symbols will refer to the corresponding
nondimensional variables, unless otherwise specified.
Gas-Phase
Equations
The gas-phase boundary layer equations are
O(pu) O(pt,)
- + - Ox
Oy
(1)
--So,
÷Sin,
(2)
M A T H E M A T I C A L MODEL
Figure 1 illustrates the system configuration, identifying the stagnation plane, the symmetry plane
(dot-and-dash), the gas-phase streamlines, and the
droplet trajectories. The shaded area represents
the high-temperature region as found in one of
the particular cases studied and discussed in the
foregoing. Air, fuel vapor, inert, hot products,
-u +N
_
+ ix \ O y J
XOT
- Qfvp + Se,
(3)
328
G. CONTINILLO and W. A. SIRIGNANO
the mass vaporization rate,
am
~-"
( dut
)
- n ~mt ~ + rn(u - Ul) ,
S e = - - n { q ÷rh[CpF(T - T s ) + L ] } ,
O
-0
..°
0..
"'""0....
(10)
where t~ is the total heat transferred to the,
droplet interior. Other symbols are identified in
the Nomenclature.
In writing the equations we have already made
the boundary layer assumptions. We also assume
that
o
, C
(9)
.,
/~
T
I~o -
(11)
To'
and that Cp, (r~), Le, and Pr are constant. If we
now multiply Eq. 2 by u and add the result to Eq.
3 we obtain
x
Fig. 1. Schematic of the problem.
° (c,r +
d- ~Vi -t- (t$iF -- Y i)Su,
6/
p = p,_--~z=,T
tin)
+ pv@
O (CpT +
(4)
(5)
where Pr is a constant of order 1. The terms
(~pT + u2/2) can be rewritten as
The expression for wi is given as
fi'i = (m)ibiB e x p ( - E a /(RT)CF~Co (3,
(6)
where M 2 ---- U 2 / ' r R T . If we restrict our study to
flame situations when the Mach number is low,
the second term in parentheses can be neglected,
and Eq. 3 can be substituted with
where a chemical reaction
bFF + boO --~ - bpP
pCpv
is taking place, and
.Yi
Ci = (m)i"
(7)
If the droplets are all represented by one group,
the expressions for the source terms are
S~ - n r n ,
(8)
where n is the droplet number density and rn is
= C-ZP
t g POy
r
#-O-y +Se -QCep
(3 t)
Equation 3' can also be derived from Eq. 3 by
neglecting the pressure work and the viscous dissipation term.
COUNTERFLOW SPRAY COMBUSTION MODELING
329
Boundary Conditions
with [15]
The system of Eqs. 1-4 is subject to the boundary
conditions
th = 2 7 r o f D f R ~ d h ln(1 + B M ) ,
Sh - - 2 + [(1 + ReSc)U3[max(1, Re)] °'°77 - 1]
y = +oc: u -----Uo+(X); T = To+; Y i = Yio+,
(17)
•F(BM),
" y = -cx~: u = Uo_(X); T = T o - ; Y i -= Y i o - .
(12)
Our case is a counterflow, for which the outer
flow is the potential flow of two impinging streams
•(Fig. 1), described by
Uo+ = K + x ;
Vo+ = - K + y ;
Y F s -- Y F ~
BM--
1 -- Y F s
(m)FXFs
YEs
(m)FXFs + (1 - X v s ) ( m ) '
1
(13)
If the reference frame is fixed to the stagnation
plane, overall momentum balance requires, for the
o<~boundary conditions, that p o _ K _ 2 __ po+K + 2,
if we neglect the dynamic pressure contribution of
the incoming droplets. Equating the static pressure at the stagnation plane for the potential field
and assuming that the stagnation pressures are the
same for the two streams, we have there, since
v = 0 at the stagnation plane,
OT
1 0 (r20T)
Ot -- at ~ ~
\
(22)
07"
3 7 r=R(t) - 47rR2otlPlC pl '
03T7 r=0 = o ;
(23)
that is
(
O =m \
- rs)
BT
Droplet Equations
CplSh
We use a hybrid Eulerian~Lagrangian formula. tion. A single-component, spherically symmetric
droplet, surrounded by a quasi-steady spherically
symmetric film is considered; the transient droplet
heat-up, vaporization, and motion is modeled• The
free-stream conditions are those of the local gas
phase surrounding the droplet. The equations are
Balance of mass
= -rr/,
- L),
(24)
(14)
BT =(1--BM)
rR3ol
(20)
(21)
Or J '
T ( r , O) = Tlo,
where
~-~
(19)
Balance of energy--conduction limit model
Po + Oo+K+ 2X2 = r(O) = P o + Po - K - 2 X 2 ,
"Po+K +2 = # o _ K _ 2.
(18)
'
X F s • -- e x p ( - - C I F / T s + C2F).
P
Uo- = K_x,
Vo- = K _ y .
(16)
(15)
~ -1,
1
4~ = C p / R u Le'
(25)
(26)
/Xru ----2 + [(1 + RePr)U3[max(1, Re)] °'°°7 - 1]
•F(Br),
L = Lboil
( Tcrit-Ts ) 0"38
k.Tcrit -- Tboil
(27)
(28)
In this study, we assume Re << 1; hence At'u =
Sh = 2.
330
G. CONTINILLO and W. A. SIRIGNANO
tions will be determined. We assume that similarity exists, leaving it to be proven at the end of the
analysis. We first nondimensionalize the equations
by introducing the following reference quantities:
Droplet Motion
The vector equation for droplet dynamics is
4R3
~,r
dVi
Or-d-f = ,rR2 ~o(V - Vt)[V - VtlCn
4
3
+ gTrR otg,
(29)
xt(O)=xlo,
t* = 1/K_;
L* =
r* =.to_
P* = Po; O* = O o - ;
Vt(O) = Vto.
(30)
For Cn we assume Stokes flow:
24
u* : ~ ;
Under the assumptions made, the state Eq. 5 becomes
24v
Cn - Re -- 2 I V - VtlR"
Q* = ~_.pT*; M* = o*L .3.
(31)
All the cases studied considered situations in
which droplet initial vector velocities match gas
velocities at the boundary: slip occurs due to gas
thermal expansion, and due to the curvature of the
flow. The deviation from Stokes's drag has been
evaluated to reach 12% in the worst case, and during a small extent of the droplet lifetime. However, the assumption of Stokes's drag must be relaxed if the droplet Reynolds number becomes of
the order 1 or larger. A convective correction factor will have to be included at least; the influence
of mass vaporization rate on the drag and a better
consideration of droplet mechanics in shear flows
will have to be addressed by subsequent studies.
Droplet Number Density
1
O = ~,
(33)
where 0 = T/T* is the nondimensional temperature, and all the quantities are now nondimensional unless explicitly specified. The nondimensional equations are
:so,
cOu +
(34)
cOu = -OC ~ x
+oL
cOy (o°u
\ cOy] + OSm,
(35)
where
For a given particle velocity field Vt, at steady
state, we can write, if particles are not generated
or destroyed:
c = po/(,o~o_r_)
V.(nVt)=O.
cO0
cOO 0 cO
u-ff~
x + V~y -- Pr Oy
(32)
SIMILARITY ANALYSIS
In principle, we now have all the equations to
solve the problem. In this section, it will be
shown that the problem is suitable for a similarity
analysis--like the simple gas-phase counterflow
flame--under certain conditions, and the condi-
(ooo
× ~, Oy,} + OSe - OQWF, (36)
uaYi
aYi = O I L
(oOYi~
Ox +o f
scoy \
+ O05iF -- Yi)S~ + Ofvi,
(37)
COUNTERFLOW SPRAY COMBUSTION MODELING
where
Yi = Yi-oo; 0 = 1
S o - o*/t*' Sm
Se
=
r/ = +c~z: f ' = -v/T+o~/T_~;
O*u*/t*'
Yi = Yi+oo; 0 = T+o~/T_o~
Se
p*CpT*/t*'
(~
Q - CpT*'
gel
fiPi
o*/t*'
and ^ denote dimensional quantities. We now introduce the similarity variable:
= f Y dy
,7
331
(38)
JoT'
f =
x d~
-
(39)
OS~d~
=-O(f+fv).
t; = Ro2/Ott;
L; = Ro;
3
M ; = -~rRo at;
Assuming similarity, all x derivatives are zero, and
from Eq. 34 we find
v =-0
has a similarity solution only if So, Sm Ix, and Se
are independent of x.
In order to nondimensionalize the droplet equations, we start with the moving boundary problem constituted of Eq. 21 with the relevant initial
and boundary conditions. The following reference
quantities are introduced:
4
and the stream function-like quantity:
(44)
T; = Tto
and the following transformation of variables is
performed:
r
f - R(t);
~fotdZ
T =
f--7'
(40)
with z a dummy variable and fs =
By using the Euler equation to express d p / d x as
a function of the outer flow velocity Uo = x, we
find the transformed momentum equation
f ' " + ( f + f v ) f " = (f,)2 _ 0 -- --,0Sin
(41)
go
OOI
( L dfs fXl OOt
Or
\ f s dr J Of
1 0 (2001)
-
f2
Of
f
~
'
(46)
where primes denote derivation in 7. Equation 36
becomes
+ ( f + fo)O' -----OSe + OQg'F
(45)
'
so that the problem transforms into
X
10Pr
R(t)
(42)
Or(f, 0) = 1,
001 ~=0
~-
= 0;
(47)
001
~
O
~=, -- 3~s'
(48)
and Eq. 37 becomes
where q has been nondimensionalized as follows:
1
"_L_y~.,+ ( f + fv)Y[ = --(~iF
Sc
--
Yi)OSo - OWl.
(43)
0
=
m
o-Ots
BT
- L
(49)
It is clear that the resulting problem, together with
the boundary conditions:
r/ = - o ~ : f = f - o ~ ; f ' = 1;
rn with respect to MT/t 7 and L with respect to
CptT 7. Equation 15, combined with Eq. 16 and
332
G. CONTINILLO and W. A. SIRIGNANO
aondimensionalized, becomes
d~s _
dr
27r M* L7 t7
3 M 7 L* t* pf__f_hDS ln(1 + BM),
(50)
where Of and D f are the nondimensional average
film density and mass diffusivity, respectively.
If the droplet velocities and accelerations are
aondimensionalized with respect to the gas refer~nce quantities, the vector equation of the droplet
motion becomes
dEXtd72 ~sl \-d-rT(d~s_ aO) ~dxt = abOV + b2~s2g,
(51)
where
_ M
7 =67r -ff/
*
L t* tt t,*, .
b
t/*
= -t*
By projecting Eq. 51 on to the x and the ~ axis
~¢e obtain
d2x,
d7"2
1 ( d~s
) dXl
~s \ dr - aO dr
not know about is (s, the instantaneous nondimensional radius of the droplet. It comes as a result
of the whole integration, as solution to Eq. 50,
which, in its turn, depends on BM, which is a
function of the local gas-phase fuel vapor concentration and surface vapor pressure, which are functions of the temperature at the drop surface, which
in turn results from the energy equation, which
is coupled to the gas-phase equations through the
boundary conditions, and so on. If we suppose
that the solution to Eq. 53, with initial conditions
that do not depend on x, is independent of x, then
all the droplets will have the same ~/ trajectory
and, since the gas-phase variables do not depend
on x, the free-stream history will be the same for
all the droplets. Hence, if the droplet initial radii
and temperature distributions are all the same, the
vaporization history will be the same, i.e., (s will
change with r (and with ~lt(r)) regardless of the x
location of the droplet. This constitutes one necessary condition for ~l(r) being independent of x,
in that it ensures that Eq. 53 is the same for all the
droplets. The other one is that initial conditions be
obviously chosen independent of x, that is,
~t(O) = ~to,
= a b O ~ x l +b2~s2gx,
(52)
7/[(0) = 7/[o,
(54)
and, since
dyt _ 0 d~lt ; dEyl
dz
-~r dr 2
_d %
dO (
and together the two conditions are sufficient.
The solution to Eq. 52, that is, the x trajectory,
is obviously dependent on x. Droplets move in x
and y, so determining the number density variation in the field. The initial conditions for Eq. 52
are the degree of freedom that we shall use in order to make the source terms for the gas-phase
independent of x. To do that, we must calculate
the number density as a function of the droplet
velocity field.
Equation 32 is also written as
)2
we have
dZm + l dO ( drtl' 2
----e
dT
O
t,-d; )
1 (d s
~s \ d r
- aO
)d.,
dr
V . (nVt) = Vt • V n + n V . V~ = 0,
~s 2
= ab[-O(f + f v ) ] + b2-ff-gy.
(53)
In Eq. 53 all the coefficients depend on variables
that are a function of ~/ only, if similarity exists
for the gas phase. The only quantity that we do
(55)
and, by definition,
Dn
II# • Vn = Dt
dn
dt
(56)
COUNTERFLOW SPRAY COMBUSTION MODELING
Integrating
To calculate V . 1"1 we first observe that
dr#
Dvt
Ovt
Ovt
dt - D~ - Vt • Vvt = ut-~x + vt-~y
333
(57)
nsvl = rloSoVlo,
or else
so
Ovt
Oy
1 dvt
vt dt
(58)
since Ovt /Ox = 0 for similarity. To calculate
Out/Ox, it is convenient to introduce the variable
nsOn[ = nosoOonlo,
that is,
nosoOo~[o
n -
s --
Os~i
(64)
Xl
(59)
Ulo
so that Eq. 52 becomes
d2s l(d =
dr 2
We are now ready to evaluate the source terms to
the gas phase equations. Starting with Eq. 41, we
have
)as
Oh f ^
x
= abO~--~s + b2~s2gx,
dftl
~mt--~- + ( ftt
OSm
~s \ dr - aO -~r
_~)~)
(65)
xp*u*/t*
(60)
By definition,
with the initial conditions
3M 7
t7 ~ = ~ '
d&l _ 47r/~03~/3~s2 d~s
s(O)
-
Xto
_
So;
s'(O)
=
1.
I11o
(61)
di
3
Since
fxt-
dxt
ds
Xto ds
XI = U l o S, Ul -- _ _
dt - U t ° d t - s d t '
Yet ds
s dt
Ou,
l dS
0 (lds~
Ox - s dt + X#° ox k s dt J "
XI
--
If we choose So to be constant, then s will not
depend on x, being a solution to Eq. 60 with initial
conditions 61 that are independent of x. Therefore
we have
L* 1 xt ds
t; ~s s d r '
ds
s dt"
1
1 dot
n d-t + ~ d t - +
=
Uto
= const,
S
dill
di
Yet d2s
s di 2
(62)
_
Note that So = Xto/Uto = 1 corresponds to matching gas and droplet velocity components along x at
the starting location for the droplets, since, for the
gas phase, in the outer flow, we have x / u = 1/K
in dimensional variables. We now have expressions for all the terms in Eq. 55, which becomes
1 dn
-
and, remembering that
we have
Out
Ox
dt
1 ds
s ~
-- 0.
(63)
L*
1
xl ['d2s
t/.2 ~s2 S ~ d r 2
1
d~sds
~s d r d r
)
and, finally, from Eq. 39 we know that
L*
~ = - - t , f 'x •
After substitutions into Eq. 65, with the aid of the
expression for n just found as Eq. 64, we finally
334
G. CONTINILLO and W. A. SIRIGNANO
obtain
OSm
- -X
M,. ( , . ) 2
-- M* \-i-it] n°s°O°Tl[°
dEs __ 3 t;
~ + ~s~
t* ~sf'S
d~s d s
2~
X
S21,1[
(66)
independent of x.
For Eq. 42 we have
-- OSe =
M*
Oh
L,3CpT*/t *
X {q -l-m[CpF(7: --'Fs)
-']-L]},
(67)
which, remembering the expression for q given
by Eq. 24, becomes
- OS~ =
OhDICpF(T M * C_.pT*
L .3
7"s)1 + B r
BT
t*
That is,
CpF M ; t* .
--OSe =
C p M't--7 m
x
T; °
1 + Br nosoOoTl:o
0 - T* Is) BT
sTq[
(68)
For Eqs. 43, we have
^ ."
r/m
OSo = Op, /t-----z.
(69)
That is,
OSo =
M~ t* nosoOo~[o 3~sd~z .
M* 7;
s~;
(70)
NUMERICAL SOLUTION PROCEDURE
The equations for the gas phase with the relevant boundary conditions constitute a boundary-
value problem, while the droplet equations, in
their lagrangian form, constitute an initial value
problem--with the exception of the liquid-phase
energy equation, which is a parabolic partial differential equation.
The gas-phase, droplet, and liquid-phase problems are coupled to each other through source
terms and/or boundary conditions. An iterative
procedure is employed. The gas-phase equations
are solved first, followed by the droplet equations
and the liquid-phase energy equation by using the
just computed values of the gas-phase variables
in the coupling terms. The source terms are then
calculated and fed back to the gas-phase equations, which are solved again, so forth. The equations for the gas phase are discretized by means of
finite-difference schemes, and a fictitious timelike
independent variable is introduced. In this way,
the equations are integrated in sequence, and the
"steady-state" solution is sought that corresponds
to the initial guess. The convergence criterion for
this inner iteration loop involving the gas-phase
equations is chosen not so severe as the overall
criterion (10 -4 relative error between two iterations), since the droplet contributions are to be
evaluated with the updated gas-phase variables,
and there is no point in calculating an accurate
solution with the incorrect source terms at each
cycle of the outer, main iteration loop.
Due to the nonlinearity of the problem, multiple
solutions exist for the same set of boundary conditions. For example, both for a diffusion and a
premixed configuration in laminar flow, the physical problem always has a stable "cold" solution
with no flame, and may or may not have stable
solutions with flames. In the affirmative case, the
choice of the initial guess, and the numerical procedure itself, determine the convergence to either
of the solutions. With the presence of droplets,
more nonlinearity is added to the problem, and
there appears to be no easy way to tell how many
different solutions the problem has. In this study,
no such analysis is attempted. When the influence
of a parameter is studied, the parameter is varied
by small increments from one case to the next,
and the last computed solution is used as a first
guess.
COUNTERFLOW SPRAY COMBUSTION MODELING
TABLE l
Main Values o f P a r a m e t e r s U s e d in the Calculations
B = 4 . 6 × 10 ]1 s - ] g - m o l -°'75
E a = 3.0 X 10 4 cal g - m o l - ~ K t~ = 0 . 2 5
8 = 1.5
Cp = 1.29 x 1 0 3 J k g - I K - I
Po = 1.0133 x l 0 s Pa
Q = 4.43504 x 107Jkg -l
p / = 7.03 x 1 0 2 k g m -3
C p / = 2 . 4 2 x 1 0 3 J k g - I K -1
L~,il = 2 . 9 7 8 x l 0 s J k g -1
Tbon = 3 . 9 8 8 X 10 z K
To = 3 x 102 K
#o = 1.8 x 10 - S k g m - I s -I
RESULTS A N D D I S C U S S I O N
Three basic configurations have been considered
in the present work. The first configuration is a
symmetric stagnation flow field, generated by two
impinging air streams at 300 K, perturbed by a
n-octane spray injected from the left side in our
representation. The second and the third are analogous to a simple gaseous counterflow premixed
2
/
/
o
o
>
LC
>"
]
.1
!
l
!
f
i
#
i -
"....
tl.
O.C
-15.0
,,,--Ii ?d..
tPt
.
!!
",.
.
.
.
,
.
!1
" .54
! ni
0.0
,
'
y [mm]
15.0
Fig. 2. Results o f calculations for the base case. Profiles
o f fuel v a p o r ( Y F , dotted) a n d o x y g e n ( Y O , broken) mass
fractions, t e m p e r a t u r e (T, solid line) a n d log o f the reaction
rate (dot a n d dash, a r b i t r a r y units). Initial droplet diameter:
D o = 50 # m . Strain rate o f the outer flow: K = 55 s - l .
335
and diffusion flame, respectively. A cold stream
containing the spray (at the left in our representation) and a hot stream form a stagnation flow.
In the former, "premixed" case, the cold stream
consists of air carrying a n-octane spray at 300
K, while the hot stream is inert at 1200 K. In the
latter, "diffusion" case, inert and droplets come
from the left at 300 K, hot air (1200 K) from the
right.
Equation 6 represents the simplest way of dealing with chemical kinetics (one-step overall reaction). The form for the reaction rate and the numerical values for the parameters are those suggested by Westbrook and Dryer [16] for a premixed flame of n-octane in air. In fact, it is likely
that this kinetic model is inadequate for accurate
predictions of flames in configurations other than
that explored by the cited authors. Moreover, limit
situations like those of flame extinction certainly
cannot be adequately described by a global, onestep kinetic model, no matter what values are used
for the parameters in the reaction rate formula. Table 1 presents the values of the kinetic parameters
along with those of other properties used in the
calculations.
In all of the cases considered, the velocities of
the droplets match the gas velocities at the left
boundary, as the temperatures (300 K) do; the
initial equivalence ratio is l; body forces are neglected; no fuel vapor is injected with gas flow.
Figure 2 shows the y profiles of fuel vapor, oxygen, temperature, and the logarithm of combustion
reaction rate for a case with K = 55 s - t and initial
droplet diameter of 50 #m, which is the base case
for the first configuration. The flame structure is
interesting in that it consists of two main reaction
zones separated by a region in which most of the
vaporization occurs. The first flame, at left, has
a predominant diffusion character, but some prevaporization occurs, as shown by the fuel vapor
profile. It appears that the droplets cross the first
flame at left and vaporize in the high-temperature
region, closer to the stagnation plane (at y = 0).
Due to both the curvature of the gas flow and
the thermal expansion, the mixture becomes rich
as the stagnation plane is approached, since the
droplets do not follow the gas motion promptly.
336
G. CONTINILLO and W. A. SIRIGNANO
~r
=o
t.-
.2
o
0
>-
u."
t.C
>.-
>,.
/
...
t
-'--~ ,.
I
",
al
1.1
;
;
I
O.Q
......
~
"../l
; l[
..~.,~
~,~
-15.0
I
15.0
0.0
0.0
-15.0
,
ir L......
.d."f
,
0.0
15.0
y [mm]
y [mm]
(a)
(c)
Fig. 3. (Continued)
=o
£_
.2
o
>,.
u."
>..
I
i [ .:. ~".,1
O.G
-15.0
I
I
,I
L_
I
0.0
I
15.0
y [mm]
Co)
Fig. 3. Influence of the strain rate. Profiles of fuel vapor and
oxygen mass fractions, temperature, and log of the reaction
rate. Initial droplet diameter: Do = 5 0 / z m . (a) K = 100 s - I .
(b) K = 3 0 0 s
- l . (c) K = 5 0 0 s
-t.
A second, pure diffusion flame stabilizes approximately at the stagnation plane.
Figure 3e illustrates the influence of the strain
rate of the flow, for a given droplet initial diameter. The strain rate is increased from that of
the base case in Fig. 2 up to 500 s -1. It appears
that increasing the strain rate causes the two react-
ing zones to narrow and approach the stagnation
plane. At the highest value, the two peaks in the
reaction rate profile are very close to each other,
while the temperature profile no longer shows two
distinct maxima. This indicates that the two reaction zones will merge into the zone at high strain
rates. Figure 4 illustrates the influence of the initial diameter of the droplets on the flame structure
for the first configuration. It is seen that, as the
droplets are chosen larger, they are able to penetrate the first flame, thus producing a peak in
the vapor fuel mass fraction, corresponding to the
droplet consumption, and higher vapor fuel mass
fraction in between the two flames.
As already noted, the particle-gas relative motion is responsible for the differences observed
with respect to the known behavior of simple
gaseous systems. Figure 5 illustrates the gas and
particle velocity components perpendicular to the
stagnation plane for the cases of Figs. 2 and 4c,
respectively. The gas velocities (solid lines) do not
differ much from the usual counterflow gas flame
profile, with a velocity jump across the flame and
the linear, potential outer flow towards the boundaries. A slight inflection can be detected corresponding to the point at which the droplets vaporize completely and cease to contribute to the
gas balance equations. It is seen that, the larger
COUNTERFLOW SPRAY COMBUSTION MODELING
337
.3
E
"t
~o
0
o
>u_
>-
LI."
>-
!
O.O
-20.0
i!
J
I
#
i
~ ....- "
0.0
~ "
I
I
20.0
0.0
-20.0
.!~.~
j
20.0
0.0
y [mm]
y [mm]
(c)
(a)
Fig. 4. ( C o n t i n u e d )
E
,i-
iii
o
.2
/
/
o> EL"
>-
!
,
I
;
\~
0.0
-20.
&-
.....,,.~.
,
=
,~d
0.0
,
,
20.0
y [rnm]
(b)
Fig. 4. Influence o f the initial droplet diameter. Profiles of
fuel vapor and oxygen mass fractions, temperature, and log of
the r e a c t i o n rate. Strain rate o f the o u t e r flow: K = 55 s - I
(a) Do = 1 0 / ~ m . (b) Do = 3 0 # m . (c) Do = tO0 #m.
are the droplets, the larger is the mismatch between the velocities of the two phases. The same
phenomenon (not shown) occurs for the velocity
component along the stagnation plane; the gas is
always accelerating and so are the droplets, which
are dragged by the gas and therefore move more
slowly.
By combining the two effects (droplet initial diameter and flow strain rate), it can be said that
higher strain rate for larger droplets produces
a more compact flame zone if compared with
smaller droplets (Fig. 6).
The second configuration (Fig. 7) differs from
the first one in that it has no oxygen in the right
stream, and therefore only one flame zone is observed. The temperature of the right, hot stream
is "chosen lower than the maximum temperature
reached in the flame; however, it appears to have
little, if any, influence on the flame. In fact, the
flame stabilizes itself out of the boundary layer,
identified by the "s"-shaped part of the temperature profile across the stagnation plane. The left
part of the structure is very similar to that found
for the first configuration (Fig. 2). This result is
obviously dependent on the particular set of the parameters chosen, first of all the strain rate; higher
strain rates will make flame and boundary layer
merge.
The third configuration (Fig. 8) produces results that are most similar to those of a pure
gaseous system. This can be explained by the fact
that, since the carrier gas does not contain oxygen, the droplets vaporize almost completely in
the preheating zone while no reaction takes place.
Different behavior can be expected for larger and
faster incoming droplets, having enough momen-
338
G. CONTINILLO and W. A. SIRIGNANO
•~
2.(1
l
r
i
"--~
E
~.,
t. . . . .
l
.2
LU
O
j
#
I-'-
O.C
~7
O
>U-- .1
>-
-2.0
-15.0
i
I
i
i
0.0
-15.0
=
0.0
15.0
y [ram]
=
'
0.0
y [ram]
15.0
Fig. 6. C o m b i n e d effect o f i n c r e a s e d droplet initial d i a m e t e r
a n d strain rate. Profiles o f fuel v a p o r a n d o x y g e n m a s s fractions, temp°erature, a n d log o f the reaction rate. Strain rate
o f the outer flow: K = 110 s - 1 . Initial droplet diameter:
Do = 100 # m .
(a)
~.o
stream. Figure 9 shows the results for K = 25
s - t , wherein the flame zone is broadened.
E
CONCLUSIONS
A similar solution for a planar counterflow spray
flame with transient heating and vaporization of
o.o
.3
,
,
,
E
-2.0
"1 5m0
i
I
I
,
0.0
",t
,
15. 0
y [ram]
Co)
Fig. 5. Influence o f the initial droplet d i a m e t e r on interphase
slip. G a s - p h a s e (solid) a n d droplet (dotted) velocity c o m p o n e n t
p e r p e n d i c u l a r to the s t a g n a t i o n plane (at y = 0). Strain rate o f
the outer flow: K = 55 s - 1 . (a) D o = 5 0 # m . (b) Do = 100
/~m.
wo .2
I--
o) . .
uS
>-
I
,1
•
tum to cross the stagnation plane and possibly the
flame before complete consumption.
The temperature of the right stream is important
in this case, as opposed to the premixed case. In
fact, a diffusion flame could not be stabilized at
the strain rate of 55 s - l with cold (300 K) right
"'"
"*'"-.....
!~
O.O
-20.0
~
"-..
..,.-.-L ; M
I
0.0
"...~
,
y [ram]
,
20.0
Fig. 7. Results o f calculations for the second configuration.
Profiles o f fuel v a p o r a n d o x y g e n m a s s fractions, t e m p e r a t u r e
a n d log o f the reaction rate. Initial droplet diameter: D o = 50
# m . Strain rate o f the outer flow: K = 55 s -~ .
COUNTERFLOW SPRAY COMBUSTION MODELING
.3
i
i
E
/
ll.-
.2
\
uS
>-
0.C
-20.,
I
I
.... ~°'°
I
I
I
0.0
20.0
y [mm]
Fig. 8. Results of calculations for the third configuration. Profiles of fuel vapor and oxygen mass fractions, temperature and
log of the reaction rate. Initial droplet diameter: Do = 50/zm.
Strain rate of the outer flow: K = 55 s -1 . Temperature of the
right stream: To+ = 1200 K.
droplets has been obtained. The mathematical formulation is complete in that it accounts for interphase exchange of mass, momentum and energy.
It is noteworthy that numerical solutions can be
computed within 10 to 30 minutes on 1 MIPS com.3
i
i
i
i
i
339
puter. This means that parametric studies can be
easily conducted.
Various configurations of counterflow of air,
droplets, and inert gas can be considered. In some
configurations, two distinct flame zones are encountered. Droplets, in general, do not follow gas
streamlines. Flames with both premixed-like and
diffusion-like characters are found. Initial droplet
size and flow strain rate are important parameters. An increase in initial droplet diameter tends
to cause a separation of the flame into a premixed
flame zone and a diffusion flame. Increase of the
strain rate tends to make the flame more compact
and to inhibit such separation.
A comparison of the results with values coming
from experimental measurements is highly desirable; however, no such a configuration of a spray
flame is available to the authors' knowledge. The
experimental setup is seen as one of the most interesting areas of development for future work.
As mentioned earlier, particle drag is to be investigated more thoroughly both with theory and
experiments. Finally, more detailed description of
the combustion chemistry will be necessary in order to make model results more realistic, especially in the study of chemically controlled situations, like, for example, flame extinction.
i
REFERENCES
1.
i'-
.2
l
>.
2.
3.
i
4.
5.
i
.1
6.
....
/
7.
..........
O.C'
-20.4
"f
~
I
I
8.
I
20.0
0.0
y [rnml
Fig. 9. Results of calculations for the third configuration. Profiles of fuel vapor and oxygen mass fractions, temperature, and
log of the reaction rate. Initial droplet diameter: Do = 50/~m.
Strain rate of the outer flow: K = 55 s -1 . Temperature of the
right stream: To+ = 300 K.
9.
10.
11.
Krishnamurthy, L., Williams, F. A., and Seshadri, K.,
Combust. Flame 26:363 (1976).
Tsuji, H., Prog. Ener. Combust. Sci. 8:93 (1982).
Libby, P. A., and Williams, F. A., Combust. Flame
44:287 (1982).
Lifian, A., Acta Astronaut. 1:1007 (1974).
Hamins, A., and Seshadri, K., Twentieth Symposium
(International) on Combustion, The Combustion Institute, Pittsburgh, 1984, p. 1903.
Seshadri, K., Puri, I., and Peters, N., Combust. Flame
61:237 (1985).
Buckmaster, J., and Mikolaitis, D., Combust. Flame
47:191 (1982).
Libby, P. A., and Williams, F. A., Combust. Sci. TechnoL 31:1 (1983).
Libby, P. A., Lifian, A., and Williams, F. A., Combust.
Sci. Technol. 34:257 (1983).
Libby, P. A., and Williams, F. A., Combust. Sci. TechnoL 37:221 (1984).
Libby, P. A., and Williams, F. A., Combust. Sci. TechnoL 54:237 (1987).
340
12.
Niioka, T., Mitani, T., and Takahashi, M., Combust.
Flame 50:89 (1983).
13. Graves, D. B., and Wendt, J. O. L., Nineteenth Symposium (International) on Combustion, The Combustion
Institute, Pittsburgh, 1982, p. 1189.
14. Puri, I., and Libby, P. A., 1988 Spring Meeting of the
Western States Section, The Combustion Institute, Paper
no. 88-42.
G. C O N T I N I L L O and W . A. S I R I G N A N O
15.
Abramzon, B., and Sirignano, W. A., Second
ASME-JSME Joint Thermal Engineering Conference,
Hawaii, March 1987.
16. Westbrook, C. K., and Dryer, F. L., Prog. Ener. Cornbust. Sci. 10:1-57 (1984).
Received 23 January 1989; revised 21 November 1989