You are on page 1of 17

Combustion and Flame 148 (2007) 100–116

www.elsevier.com/locate/combustflame

Extinction of counterflow diffusion flames with radiative


heat loss and nonunity Lewis numbers
H.Y. Wang a , W.H. Chen b , C.K. Law a,∗
a Department of Mechanical and Aerospace Engineering, Princeton University, Princeton, NJ 08544-5263, USA
b Department of Marine Engineering, National Taiwan Ocean University, Keelung, Taiwan, Republic of China

Received 12 July 2006; received in revised form 30 October 2006; accepted 31 October 2006
Available online 12 December 2006

Abstract
The structure and extinction characteristics of counterflow diffusion flames with flame radiation and nonunity
Lewis numbers of the fuel and oxidant are examined using multiscale asymptotic theory, and a model expressed in
terms of the jump relations and reactant leakages with the proper consideration of the excess enthalpy overlooked
in previous analyses is developed. The existence of the dual extinction limits in the presence of radiative heat loss,
namely the kinetic limit at small Damköhler number (high stretch rate) and the radiative limit at large Damköhler
number (low stretch rate), are identified. It is found that the former is minimally affected by radiative loss, while
a substantial amount of heat loss is associated with the radiative limit. Reactant leakage, however, is the root
cause for both limits. The influence of radiative loss on the extinction Damköhler numbers is found to be through
its effects on the flame temperature, the excess enthalpy, and the reduced extinction Damköhler number. At both
extinction limits, the contribution from the flame temperature is always important and dominant. The contributions
from the other two, however, could be important in some special cases. At small LeF , the contribution from the
reduced extinction Damköhler number is large and even dominant under small radiative loss. The contribution
from the excess enthalpy is important for small LeO and it may be comparable to the contribution from the flame
temperature when radiative loss is small. Thus, overlooking the excess enthalpy in previous analyses may have
resulted in rather large error in the predicted extinction Damköhler numbers, especially the kinetic one.
© 2006 The Combustion Institute. Published by Elsevier Inc. All rights reserved.

Keywords: Diffusion flame; Excess enthalpy; Extinction; Asymptotics

1. Introduction It is shown that the flame structure and response are


controlled by the system Damköhler number, Da, de-
The structure and extinction characteristics of dif- fined as the ratio of the residence time to the reaction
fusion flames in the absence of volumetric heat loss time, and that there exists a minimum Da, and conse-
were first rigorously analyzed using large-activation- quently a maximum flow rate, beyond which steady
energy asymptotics in a seminal paper by Liñán [1]. burning is not possible due to insufficient time for
adequate reaction to occur to maintain the flame struc-
ture. This extinction mechanism is intrinsic to the
* Corresponding author. Fax: +1 609 258 6233. structure of diffusion flames, and hence is operative
E-mail address: cklaw@princeton.edu (C.K. Law). under all situations.

0010-2180/$ – see front matter © 2006 The Combustion Institute. Published by Elsevier Inc. All rights reserved.
doi:10.1016/j.combustflame.2006.10.005
H.Y. Wang et al. / Combustion and Flame 148 (2007) 100–116 101

Nomenclature

BK Pre-exponential factor for the reaction Δ Reduced Damköhler number


rate Δc Minimum value of the reduced Damköh-
cp Specific heat ler number
Da Damköhler number ε Reciprocal of the Zel’dovich number
DaC Collision Damköhler number η Stretched spatial coordinate in the reac-
DaE Extinction Damköhler number tion zone
hi Excess/deficiency enthalpy for species i, κ Planck’s mean absorption coefficient
i = F, O ν Stoichiometric mass ratio of oxidant to
htotal Total excess/deficiency enthalpy in the fuel
reaction zone νi Stoichiometric coefficient of species i,
k Stretch rate of flow i = F, O
Lei Lewis number of species i, i = F, O σ Stefan–Boltzmann constant
q Heat of combustion per unit mass of fuel ω Chemical reaction rate
supplied
ζ Stretched spatial coordinate in the radia-
qR Rate of radiative heat loss per unit vol-
tion zone
ume
Ra Radiative heat loss parameter Overbars, subscripts, and superscripts
Si Leakage function of species i, i = F, O ∼ Unscaled quantities for T and Yi ,
xf Flame sheet location i =F, O
T Temperature ∗ Value for reference state
Tf Flame temperature
−∞ Fuel boundary
Ta Activation temperature
∞ Oxidant boundary
Wi Molecular weight of species i, i = F, O
+ Oxidant side of flame sheet
Yi Mass fraction of species i, i = F, O
− Fuel side of flame sheet
Greek symbols ad Adiabatic state
Γ Relative strength of the radiative loss to ext Extinction condition
the chemical heat release F Fuel
Γc Maximum radiative strength a system K Extinction condition at the kinetic limit
can sustain O Oxidant
γ Heat transfer parameter R Extinction condition at the radiative
δ Order of radiation zone thickness limit

Extinction can be further promoted in the presence tion is expected to be facilitated with increasing Da.
of radiative heat loss [2–4]. Specifically, Sohrab et al. Consequently, in addition to the extinction limit at the
[4] analyzed the structure and extinction of counter- minimum Da, there should exist another extinction
flow diffusion flames with flame radiation, recogniz- limit at a maximum Da, above which steady burning
ing that radiation is a temperature-sensitive process, is also not possible.
albeit less sensitive than chemical reaction. Conse- The possible existence of the dual extinction lim-
quently, radiative loss is operative within a thin, O(δ), its at lower and higher Damköhler numbers, hereafter
zone that sandwiches the O(ε) reaction zone, but respectively referred to as the kinetic and radiative ex-
is nevertheless embedded within the much thicker tinction limits, was first suggested and numerically
O(1) outer diffusive–convective zone, where ε is the demonstrated by T’ien [5] for the counterflow dif-
reciprocal of the Zel’dovich number, and δ satisfies fusion flame in the stagnation region of a condensed
ε  δ  1. Multiscale asymptotics using δ and ε as fuel with surface radiation. Theoretically, Chao et al.
small parameters showed that the radiative contribu- [6] first successfully demonstrated, via the multiscale
tion to extinction is through the reduction of the flame asymptotic theory of Sohrab et al. [4], the existence
temperature. However, an important qualitative ex- of dual extinction states for droplet combustion with
tinction behavior was not recognized. That is, because flame radiation. The analysis further showed that ex-
the extent of radiative loss increases with increas- tinction for both limits is governed by Liñán’s ex-
ing flame volume and thereby flame thickness, and tinction criterion as a consequence of excessive re-
because flame thickness increases with decreasing actant leakage. Subsequently, this multiscale asymp-
stretch rate and hence increasing system Da, extinc- totics was employed by Oh et al. [7] for the diffusion
102 H.Y. Wang et al. / Combustion and Flame 148 (2007) 100–116

flame stabilized on a condensed fuel with both flame a model for distinct and general Lewis numbers of
and surface radiation and by Liu et al. [8] for the fuel and oxidant, LeF and LeO , and with proper con-
counterflow diffusion flame with flame radiation and sideration of the excess enthalpy. This formulation is
small deviation of the Lewis numbers of fuel and ox- then applied to study the dual extinction limits of the
idant from unity. Specifically, Liu et al. [8] showed radiative counterflow diffusion flames with nonunity
that larger Lewis numbers lead to a smaller flammable Lewis numbers. It should be mentioned that Mills
range of Damköhler number and smaller Lewis num- and Matalon [11] studied the extinction of burner-
bers increase this flammable range. Experimentally, generated radiative spherical diffusion flames with
the existence of the dual extinction limits was ob- nonunity LeF and LeO . However, the radiative loss
served by Maruta et al. [9] in a counterflow diffusion was assumed to be of O(ε) quantity and occurs within
flame of methane and air with flame radiation under a zone of O(1) thickness between the burner surface
microgravity. and flame, whose location and thickness are arbitrary
The theoretical analyses mentioned above, how- given. Thus, this analysis is based on different as-
ever, did not consider the excess/deficiency of the sumptions from the current study, in which radiative
total enthalpy in the reaction zone, hereafter referred loss occurs within the radiation zone of O(δ) thick-
to as excess enthalpy. Its importance was recognized ness determined from the temperature sensitivity of
by Kim and Williams [10] in their study of coun- radiation, and hence has a different scope of applica-
terflow diffusion flame with nonunity Lewis number, tion. Furthermore, Mills and Matalon [11] found that
Le. The excess enthalpy arises from the nonconserva- the Lewis numbers of fuel and oxidant have a signifi-
tive nature of the total enthalpy due to the imbalance cant effect on the extinction limits, and steady burning
of thermal and mass diffusion as the reactants leak is not possible when they are sufficiently large. This
through the flame as a result of finite-rate chemistry. substantiates the need to conduct a rigorous extinc-
It was shown that, although the amount of the ex- tion analysis for Lewis numbers sufficiently different
cess enthalpy is small, typically of O(ε), it can lead from unity.
to O(1) changes in the reaction rate. Furthermore, We further note that since the formulation devel-
without considering the excess enthalpy, the depen- oped in the current study is of a general nature, it
dence of Da on the reduced Damköhler number, Δ, is can be applied to other phenomena affected by simul-
linear [1], so that the minimum of Da directly cor- taneous radiative loss and mixture nonequidiffusion,
responds to the minimum of Δ, Δc . The physical such as the thermal–diffusive instability [12–18] of
extinction limit can be consequently determined from radiation-affected diffusion flames. This possible ex-
the reaction-sheet solution using Liñán’s formula [1]. tension will be discussed subsequently.
However, with the consideration of excess enthalpy,
this dependence becomes nonlinear, so that Δc does
2. Formulation
not necessarily correspond to the minimum of Da,
and as such does not correctly identify the extinction 2.1. Governing equations
condition. Recognizing that in addition to nonequid-
iffusion, which is the cause of excess enthalpy in the Fig. 1 shows the counterflow configuration consid-
analysis of Kim and Williams [10], radiative loss from ered in this study, with the fuel and oxidant streams
flames is the other major source of enthalpy loss in approaching from the left- (−∞) and right-hand (∞)
flames, overlooking the excess enthalpy is expected to sides, respectively. The chemical reaction is assumed
produce O(1) error in the radiative extinction limit, at to be described by a one-step irreversible reaction of
which radiative loss is significant. Furthermore, pre- the form
vious analyses [4,6–8] did not consider adequately
νF F + νO O → products,
the effect of nonequidiffusion. Since extinction can
be considered to be primarily attributable to exces-
sive heat loss from the reaction zone, which could be
achieved through both conduction and radiation, it is
necessary to consider the effects of nonunity Le and
radiative loss at the same time. Although Liu et al.
[8] considered nonunity Le with O(ε) deviation from
unity, such a small deviation cannot explore all the Le
dependency of the dual extinction limits, as will be
shown later.
In view of the above considerations, one of the pri-
mary objectives of the current study is to perform a
rigorous multiscale asymptotic analysis and develop Fig. 1. Schematic of a counterflow diffusion flame.
H.Y. Wang et al. / Combustion and Flame 148 (2007) 100–116 103
 
T − Tf
where F and O represent the fuel and oxidant, respec- qR = qRf exp , (5)
tively, and νi (i = F, O) the corresponding stoichio- δTf
metric coefficients. Further assuming constant physi- where qRf = 4σ κq 3 Tf4 /(ρcp4 k) and δ = d(ln T )/
cal and chemical properties of the reactants, constant d(ln qR ). With the assumptions of constant density
density, and an optically thin approximation for the and properties made in this study, δ = 0.25.
radiative loss, the appropriate nondimensional gov-
erning equations can be written as 2.2. Asymptotic expansions

∂T ∂T ∂ 2T Following Chao et al. [6], the temperatures T


−x − = ω − qR , (1)
∂t ∂x ∂x 2 in the diffusive–convective, radiation, and reaction
∂YF ∂YF ∂ 2 YF zones are respectively expanded in terms of ε and δ as
−x − Le−1 = −ω, (2) ±
∼ T0± (x, t) + εT1± (x, t)
∂t ∂x F ∂x 2 Tout
∂ 2 YO  
∂YO ∂Y
− x O − Le−1 + δ T2± (x, t) + εT3± (x, t) + O(δ 2 ),
O ∂x 2 = −ω, (3) (6)
∂t ∂x  
TR± ∼ Tf + εΘ1± (ζ ) + δ Θ2± (ζ ) + εΘ3± (ζ )
with the boundary conditions
+ O(δ 2 ), (7)
T = T−∞ , YF = YF,−∞ , YO = 0,
Tin ∼ Tf + ετ1 (η) + ε2 τ2 (η) + O(ε3 ), (8)
as x → −∞,
where ζ = (x − xf )/δ and η = (x − xf )/ε are the
T = T∞ , YF = 0, YO = YO,∞ ,
stretched spatial coordinates in the radiation zone and
as x → ∞, (4) reaction zone, respectively, and xf is the flame sheet
where T is the temperature, Yi the mass fraction of location. The superscripts “+” and “−” denote solu-
species i (i = F, O) and Lei the corresponding Lewis tions on the oxidant and fuel sides of the flame, re-
number, ω = DaC YF YO exp(−Ta /T ) the chemical re- spectively. The reason for expanding Ti only in terms
action rate, qR = 4σ κq 3 T 4 /(ρcp4 k) the rate of radia- of ε is that the reaction zone is very thin so that the
tive heat loss, DaC = BK ρYF,−∞ νO /(WF k) the colli- radiative loss from it can be assumed to be negligible.
Because YF and YO are not directly affected by
sion Damköhler number, and Ta , k, σ , κ, q, ρ, cp , BK ,
radiative loss, their analyses only need to be con-
Wi , respectively denote the activation temperature,
ducted in two zones, namely the reaction zone and
stretch rate of the flow, Stefan–Boltzmann constant
the diffusive–convective zone. Thus, they can be re-
(1.36 × 10−11 kcal/m2 s K4 ), Planck’s mean absorp-
spectively expanded as
tion coefficient, heat of combustion per unit mass of
± ± ± ±
fuel supplied, density, specific heat, pre-exponential Yi,out ∼ Yi,0 (x, t) + εYi,1 (x, t) + ε2 Yi,2 (x, t)
factor for the reaction rate, and molecular weight of
species i. The nondimensional T and Yi are defined as + O(ε3 ), (9)
Yi,in ∼ εyi,1 (η) + ε2 yi,2 (η) + O(ε3 ), (10)

T= , YF = ỸF , YO = ỸO /ν, where i = F, O. We note that unlike the analysis of
(q/cp )
[6], the outer expansions for YF and YO do not have
where ν = νO WO /νF WF is the stoichiometric mass the O(δ) terms. This is because these terms do not
ratio of the oxidant to the fuel and the over-tilde “∼” have the corresponding matching terms in the reac-
designates the unscaled T and Yi for differentiation tion zone and because of their homogeneous bound-
with their nondimensional counterparts. ary conditions at x → ±∞. Consequently, it can be
With the assumption of large activation energy, shown that they are identically zero throughout the
chemical reaction is confined within a thin zone of diffusive–convective zone.
O(ε) thickness, where ε = Tf2 /Ta and Tf is the flame The above expansions for T are subject to the
temperature. Due to the temperature-sensitive nature matching conditions between the radiation and reac-
of radiation, radiative loss is assumed to be confined tion zones,

within a radiation zone of O(δ) thickness, with the dΘ1± (0) ⎪

reaction zone embedded within it, as shown by the Θ1± (0) ∼ τ1 |η→±∞ , ∼ 0, ⎪

dζ ⎪

flame structure in Fig. 1. Further recognizing that ±  ⎪

dΘ (0) dτ  ⎬
the diffusive–convective zone external to the radiation ± 2 1
Θ2 (0) ∼ 0, ∼  ,
zone is of O(1) thickness, we seek a solution for the dζ dη η→±∞ ⎪ ⎪
 ⎪

flame structure satisfying the relation ε  δ < 1. Fol- ±  ⎪

± dΘ (0) dτ ⎪
lowing Sohrab et al. [4], the rate of radiative loss, qR , Θ3 (0) ∼ 0, 3
∼ 2
 ,⎪⎭
dζ dη η→±∞
is approximated by an Arrhenius-type function,
(11)
104 H.Y. Wang et al. / Combustion and Flame 148 (2007) 100–116

and the matching conditions between the diffusive– where Da = ε3 DaC exp(−Ta /Tf ) is the reaction
convective and radiation zones, Damköhler number. Integrating Eq. (16) twice and
⎫ applying the relevant matching conditions, (11) and
Tf ∼ T0± (xf ), ⎪
⎪ (13), for τ1 , yi,1 , dτ1 /dη, and dyi,1 /dη yield


± ±
Θ1 |ζ →±∞ ∼ T1 (xf ), ⎪




⎪ dΘ2+ (0)
±  ⎪ 1 1 +
dΘ1



⎪ τ1 + yF,1 = η + Θ1+ (0) + Y ,
 ∼ 0, ⎪
⎪ LeF dζ LeF F,1
dζ ζ →±∞ ⎪


⎪ (18)
± ⎪

Θ2 |ζ →±∞ ∼ −∞, 1 dΘ2− (0) 1 −
 (12) τ1 + yO,1 = η + Θ1− (0) +
∂T0± (xf ) ⎪
Y
dΘ2±  ⎪
⎪ LeO dζ LeO O,1
 ∼ ,⎪⎪
dζ ζ →±∞ ⎪
⎪ (19)
∂x ⎪



±
Θ3 |ζ →±∞ ∼ −∞, ⎪
⎪ and the jump relations




± 
dΘ3 ∂T1 (xf ) ⎪
± ⎪
⎪ dΘ2+ (0) dΘ2− (0)

 ∼ ⎪ ∂YF,0

dζ ζ →±∞ ∂x
. ⎭ − = −Le−1
F
dζ dζ ∂x

The expansions for YF and YO are subject only ∂YO,0
= −Le−1
O , (20)
to the matching conditions between the diffusive– ∂x
convective and reaction zones:
⎫ Θ1+ (0) − Θ1− (0) = −Le−1 −1
F [YF,1 ] = −LeO [YO,1 ].
±
0 ∼ Yi,0 (xf ), ⎪
⎪ (21)


±
∂Yi,0 (xf ) ⎪

± ⎪
⎪ Integrating Eq. (17) once and applying the relevant
yi,1 |η→±∞ ∼ η + Yi,1 (xf ), ⎪

∂x ⎪
⎪ matching conditions, (11) and (13), for dτ2 /dη and
 ⎬
± dyi,2 /dη yield the jump relation
dyi,1  ∂Y (x
i,0 f ) (13)
∼ , ⎪

dη η→±∞ ∂x ⎪



dΘ3+ (0)
 
dΘ3− (0)


 ⎪ xf Θ1+ (0) + − xf Θ1− (0) +
dyi,2 
±
∂ 2 Yi,0 (xf ) ±
∂Yi,1 (xf ) ⎪


⎪ dζ dζ
 ∼ 2
η+ .⎪

dη η→±∞ ∂x ∂x ∂YF,1
= − xf YF,1 + Le−1
F ∂x
2.3. Jump relations and reactant leakages across the
∂YO,1
reaction zone = − xf YO,1 + Le−1
O . (22)
∂x
We aim to derive the jump relations and the re- So far we have derived the jump relations (14) and
actant leakages that serve as the inner boundary (20)–(22) across the reaction zone. However, for the
conditions for the outer solutions in the diffusive– current multiscale analysis, the reaction zone is sand-
convective zone. wiched into the radiation zone so that the temperature
The matching conditions for the O(1) outer solu- terms in these jump relations are expressed by the ex-
tions in Eqs. (12) and (13) yield the jump relations pansion terms in the radiation zone. Since we aim to
+ − derive the jump relations that are to be used as the
[T0 ] = [YF,0 ] = [YO,0 ] = YF,0 (0) = YO,0 (0) = 0, inner boundary conditions of the outer expansions,
(14) the terms Θ1± (0), dΘ2± (0)/dζ , and dΘ3± (0)/dζ in
where we have adopted the notation [T0 ] = T0+ (xf ) − Eqs. (20)–(22) need to be replaced by the outer expan-
T0− (xf ). Additional jump relations and reactant leak- sion terms. This can be realized through integrating
ages are to be derived through the asymptotic analysis the structure equation of the radiation zone and ap-
of the reaction zone. Substituting the inner expansions plying the matching conditions between the radiation
(8) and (10) and the stretched coordinate η into the zone and the diffusive–convective zone. Substituting
governing equations (1)–(3) yields, to O(ε), Eq. (7) and the stretched coordinate ζ into the chemi-
cally frozen form of Eq. (1) yields
d 2 τ1 /dη2 = −DayF,1 yO,1 eτ1 , (15)
d 2 Θ2± /dζ 2 = Ra exp(Θ2± /Tf ), (23)
d 2 (τ1 + yi,1 /Lei )/dη2 = 0, i = F, O, (16)
d 2 Θ1± /dζ 2 = RaΘ1± exp(Θ2± /Tf )/Tf , (24)
and to O(ε2 ),
d 2 Θ3± /dζ 2 + xf dΘ1± /dζ = RaΘ3± exp(Θ2± /Tf )/Tf ,
d 2 (τ2 + yi,2 /Lei )/dη2 = 0, i = F, O, (17) (25)
H.Y. Wang et al. / Combustion and Flame 148 (2007) 100–116 105

where Ra = δqRf . Integrating Eqs. (23)–(25) and ap- hF = A+ T1+ (xf ) + Le−1 +
F YF,1 (xf ),
plying the relevant matching conditions (see Appen-
dix A) yield hO = A− T1− (xf ) + Le−1 −
O YO,1 (xf ) (35)
are the excess/deficiency in the fuel and oxidant en-
dΘ2± (0) ∂T ± (xf )
= A± 0 , (26) thalpies, respectively, evaluated at the reaction sheet.
dζ ∂x This process is standard [1,17] and is not repeated
Θ1± (0) = A± T1± (xf ), (27) here.
  The leakage functions SF and SO are dependent
± ±
dΘ3 (0) 1 1 1 ∂T (xf ) only on Δ and γ . Liñán [1] has shown that for a given
∼ A± − ± T1± (xf ) + ± 1 ,
dζ 2 A A ∂x γ there exists a minimum Δ, Δc , given by
(28)    2  3
Δc = e 1 − |γ | − 1 − |γ | + 0.26 1 − |γ |
where A± = 1 + 2RaTf /(∂T0± (xf )/∂x)2 . Thus,  4 
+ 0.055 1 − |γ | , (36)
the jump relations in terms of the outer solutions
can be obtained by substituting Eqs. (26)–(28) into such that there is no solution for Δ < Δc and there are
Eqs. (20)–(22), respectively, two branches of solutions for Δ > Δc . Their approx-
imate formulas are given by Cheatham and Matalon
∂T ∂YF,0 ∂YO,0
A 0 = −Le−1 = −Le−1
[17] as
,
∂x F ∂x O ∂x   
d0 Δ−4/3 exp −d1 (Δ − Δc )d2 ,
(29) S1 =  
[AT1 ] = −Le−1 −1 Δ−1/3 q0 + q1 (Δ − Δc )q2 ,
F [YF,1 ] = −LeO [YO,1 ], (30)   
e0 Δ−4/3 exp −e1 (Δ − Δc )e2 ,
1 ∂T1 ∂Y
= − xf YF,1 + Le−1 S2 =  
F,1
Bxf T1 +
A ∂x F ∂x Δ−1/3 r0 + r1 (Δ − Δc )r2 ,

∂YO,1
= − xf YO,1 + Le−1
where the upper and lower expressions correspond
O , (31)
∂x to the solutions of the lower and upper branches, re-
spectively (see, for example, Fig. 2 in Ref. [14]). For
where B ± = [(A± )2 + 1]/2A± and we have used the γ > 0 and < 0, S1 and S2 correspond to (SF , SO ) and
notation [AT1 ] = A+ T1+ (xf ) − A− T1− (xf ). (SO , SF ), respectively. The coefficients di , ei , qi , and
The jump relations (14) and (29) provide enough ri (i = 0, 1, 2) only depend on γ and are given in the
inner boundary conditions for the O(1) outer so- appendix of Ref. [14]. The heat transfer parameter, γ ,
lutions. However, two additional conditions for the measures the degree of asymmetry of the thermal dif-
O(ε) outer solutions are required to close the prob- fusion across the reaction zone. If γ = 0, the tempera-
lem. They can be supplied by the amounts of leakage ture profile is symmetric and hence heat is conducted
of the fuel and oxidant through the reaction zone, and to the fuel and oxidant sides equally so that SF = SO .
are given as As γ > 0, more heat is conducted to the oxidant side
+ and consequently the oxidant is more completely con-
YF,1 (xf ) = LeF SF (γ , Δ), sumed, so that SF > SO . Similarly, SF < SO as γ < 0.

YO,1 (xf ) = LeO SO (γ , Δ) (32) The term htotal = (1 + γ )hO /2 + (1 − γ )hF /2 in
Eq. (34) represents the excess/deficiency of the to-
by transforming the energy equation (15) into the tal enthalpy in the reaction zone. This is the term
Liñán canonical form, where that was overlooked in previous analyses [4,6–8].
  Without considering it, Δ depends linearly on Da
∂T + (xf ) ∂T − (xf ) ∂T −1
γ = − A+ 0 + A− 0 A 0 so that its minimum value, DaE , coincides with that
∂x ∂x ∂x of Δ, Δc . Because γ is determined from the lead-
(33) ing order solutions, the extinction condition, repre-
and sented by DaE , can be accordingly determined di-
rectly from Δc . However, when the excess enthalpy is
∂T −2 taken into consideration, Eq. (34) becomes a nonlin-
Δ = 4LeF LeO Da A 0
∂x ear relation between Δ and Da because htotal depends
  on the reactant leakages SF and SO , which, in turn,
1+γ 1−γ
× exp hO + hF (34) depend on Δ. Consequently, Δc does not correctly
2 2
identify the extinction condition, DaE , and the deter-
is the reduced Damköhler number, SF and SO are mination of Da from a given Δ requires the O(ε)
the leakage functions of the fuel and oxidant, respec- outer solutions, which substantially complicates the
tively, and solution procedure.
106 H.Y. Wang et al. / Combustion and Flame 148 (2007) 100–116

2.4. Summary of the model and the flame location xf and temperature Tf can be
respectively determined from the jump relation (38)
So far we have derived the jump relations in implicitly as
terms of the outer solutions and the reactant leakages
across the reaction zone, which provide sufficient in- Le−1 2
F YF,−∞ exp(−LeF xf /2)
ner boundary conditions to fully determine the O(1) I − (xf ; LeF )
and O(ε) outer solutions and the flame sheet location.
They are summarized as follows: Le−1 2
O YO,∞ exp(−LeO xf /2)
=− , (45)
To O(1): I + (xf ; LeO )
 
 +   − 
+
[T0 ] = [YF,0 ] = [YO,0 ] = YF,0 −
(0) = YO,0 (0) = 0, ∂T0 (xf ) 2 ∂T0 (xf ) 2
+ 2RaTf + + 2RaTf
(37) ∂x ∂x

A
∂T0
= −Le−1
∂Y F,0
= −Le−1
∂YO,0
. Le−1 2
F YF,−∞ exp(−LeF xf /2)
∂x F ∂x O ∂x = . (46)
I − (xf ; LeF )
(38)
To O(ε): Equation (46) indicates that the chemical heat release
is conducted away from the flame to both sides. It is
[AT1 ] = −Le−1 −1
F [YF,1 ] = −LeO [YO,1 ], (39) seen that in addition to being used to heat up the re-
actants, the heat release now needs to compensate for
1 ∂T1 ∂Y
= − xf YF,1 + Le−1
F,1
Bxf T1 + F
the radiative loss through the term 2RaTf in Eq. (46).
A ∂x ∂x However, the temperature gradient decreases at the

∂YO,1 same time, due to the reduction of the flame tem-
= − xf YO,1 + Le−1
O , (40)
∂x perature through radiative loss. Moreover, it is noted
+ − that the total heat release is controlled by the reac-
YF,1 (xf ) = LeF SF (γ , Δ), YO,1 (xf ) = LeO SO (γ , Δ).
tants consumption and hence is fixed. In most cases,
(41)
these two opposite effects on heat transfer by radia-
It is noted that we have not derived the jump re- tive loss are not equal. Therefore, the overall outcome
lations for the O(δ) expansion terms of T . This is of radiative loss is to redistribute the proportions of
because the matching conditions Θ2± (0) ∼ 0 imply heat transfer from the flame to both sides, and as such
that the O(δ) terms can be solved separately on ei- it plays a role similar to that of varying the thermal
ther side of the reaction zone. diffusivities of the reactants.
Applying the homogeneous boundary conditions
2.5. Solutions of counterflow diffusion flame at x → ±∞ and the jump and leakage conditions
(39)–(41), the O(ε) outer solutions, T1± , YF,1 ±
, and
We now apply the model (37)–(41) to the coun- ±
YO,1 , can be fully determined. Here we only show
terflow diffusion flame with nonunity Lewis numbers
them in the form of the excess enthalpies, hF and hO ,
and radiative loss. The O(1) outer solutions subject
because they are the reason the O(ε) outer solutions
to the boundary condition (4) and jump relations (37)
are required. Thus we have
and (38) can be solved as

⎪ T + (Tf − T−∞ )I − (x; 1)/I − (xf ; 1), hF = CF,1 SF (γ , Δ) + CO,1 SO (γ , Δ),
⎨ −∞

x < xf ,
T0 (x) = hO = CF,2 SF (γ , Δ) + CO,2 SO (γ , Δ)

⎪ T∞ + (Tf − T∞ )I + (x; 1)/I + (xf ; 1),

x > xf , and the total excess enthalpy is then given as
⎧ (42)
 
⎨ YF,−∞ 1 − I − (x; LeF )/I − (xf ; LeF ) ,
htotal = CF SF (γ , Δ) + CO SO (γ , Δ),
YF,0 (x) = x < xf ,

0, x > xf , where
(43)
 0, x < x ,
 f  Ci = (1 − γ )Ci,1 /2 + (1 + γ )Ci,2 /2, i = F, O,
YO,0 (x) = YO,∞ 1 − I + (x; LeO )/I + (xf ; LeO ) ,  +    
x > xf , MT MT− −1 + MT+
CF,1 = − − Le M
F F −
(44) A+ A− A+
where  −1 MT−  + LeF +  
LeF A− − MF− MO − Le MF
    + O
,
I ± (x; L) = π/2L ∓1 + erf L/2x −1 + −1 −
−LeF MO + LeO MF
H.Y. Wang et al. / Combustion and Flame 148 (2007) 100–116 107

    
MT+ MT− −1 1 1 ∂T −2
CO,1 = − × exp Ta − A 0 .
A+ A− Tref Tf ∂x
  −1 M −  + −  (47)
LeF A−T − MF− (MO − MO )
× −1 + −1 −
, The extinction Damköhler number Da∗E corre-
−LeF MO + LeO MF sponds to the minimum value of Da∗ ; i.e.,
 +    
MT MT− −1 MT+
CF,2 = − − Le M
F F
+
− ∂SF (γ , Δ)/∂Da∗ → ∞. (48)
A+ A− A+
 +
  We have known that for a fixed system with given
−1 MT
LeF A+ − MF − + LeF + 
MO − Le MF Ra∗ , all the terms in Eq. (47) other than Δ and Da∗
O
+ , can be determined. Thus, Eq. (47) gives a definite re-
−1 + −1 − lation between Da∗ and Δ. Substituting Eq. (47) into
−LeF MO + LeO MF
 +   +  Eq. (48) yields the extinction condition
MT MT− −1 MT MT−  
CO,2 = − − − −
A+ A A+ A ∂h
Δ total −1=0 (49)
∂Δ ext
 MT+ − − 
Le−1 +
F A+ − MF (MO − MO ) from which the reduced Damköhler number at ex-
+
−Le−1 + −1 −
F MO + LeO MF
tinction, Δext , can be solved. Then the extinction
Damköhler number Da∗E can be solved by substitut-
and ing Δext into Eq. (47).
1 exp(−xf2 /2)
MT± = B ± xf + ,
A± I ± (xf ; 1)
3. Results and discussion
exp(−Lei xf2 /2)
Mi± = xf + Le−1 , i = F, O.
i I ± (xf ; Lei ) We use a CH4 /air counterflow diffusion flame to
We note that the total excess enthalpy htotal is a demonstrate the dual extinction characteristics with
linear combination of the reactant leakages, SF and the effects of radiative loss and nonunity Lewis num-
SO , while γ and the coefficients CF and CO are only bers. The system parameters adopted are T̃−∞ =
dependent on the leading order solutions and radia- T̃∞ = 300 K, T̃a = 24,000 K, cp = 0.334 kcal/kg K,
tive loss. Thus, for a fixed system with given radiative and q = 11,990 kcal/kg. The boundary conditions for
loss, the total enthalpy htotal is solely determined by the fuel and oxidant fractions are fixed at ỸF,−∞ = 1
the value of Δ. and ỸO,∞ = 0.23, respectively, throughout this study
unless otherwise specified. Thus, the initial mixture
2.6. Extinction analysis strength, defined as φ = ν ỸF,−∞ /ỸO,∞ , is 17.4 in
the current study.
Since the quantifying parameters ε, Ra, and Da are Fig. 2 shows variations of the flame location, xf ,
functions of the flame temperature, Tf , which varies with the fuel and oxidant Lewis numbers, LeF and
with radiative loss, it is necessary to rescale all the pa- LeO , respectively. It is seen that the flame is always
rameters to a fixed, absolute reference state [6]. Using
the adiabatic flame temperature Tref for ỸF,−∞ = 1,
ỸO,∞ = 0.23, and LeF = LeO = 1 as the reference
state, and denoting the associated quantities by the su-
perscript “∗,” we have
 2 3
Da∗ = Tref /Ta DaC exp(−Ta /Tref ),
4 ρc4 k ,
Ra∗ = 4δσ κq 3 Tref p
so that
 
Da = Da∗ (Tf /Tref )6 exp Ta (1/Tref − 1/Tf ) ,
Ra = Ra∗ (Tf /Tref )4
and Eq. (34) becomes
   Fig. 2. Variations of the flame sheet location, xf , with the
Tf 6
Δe−htotal = Da∗ 4LeF LeO fuel and oxidant Lewis numbers, LeF (LeO = 1) and LeO
Tref (LeF = 1).
108 H.Y. Wang et al. / Combustion and Flame 148 (2007) 100–116

(a)
(a)

(b)
(b)
Fig. 3. Variations of the coefficient of fuel leakage in the total
excess enthalpy, CF , with (a) LeF (LeO = 1), and (b) LeO Fig. 4. Fuel leakage SF as a function of the reduced Damköh-
(LeF = 1), for different values of radiative loss, Ra∗ . ler number Δ and Damköhler number Da∗ for LeF = LeO
= 1 and different values of radiative loss, Ra∗ .
located on the oxidant side of the stagnation surface
over the entire ranges of LeF and LeO , which is a and Ra∗ and for LeF > 1, LeO > 1, and Ra∗ > 0,
consequence of the stoichiometric nature of diffusion CF is always negative. Thus, the excess enthalpy is
flames. Furthermore, it is seen that with the decrease always negative under these parameters. For flames
of LeF (LeO ), the flame moves toward the oxidant with radiative loss, the only possibility for the excess
(fuel) side and LeF has a much stronger effect on the enthalpy to be positive is that the Lewis numbers are
flame location than LeO . This is because for the cur- sufficiently smaller than unity.
rent problem the flame is located on the oxidant side Fig. 4 shows the fuel leakage SF as a function
of the stagnation surface. Thus, fuel diffusion has to of the reduced Damköhler number Δ (Fig. 4a) and
overcome convection in the opposite direction to sup- Da∗ (Fig. 4b) for LeF = LeO = 1 and different val-
ply the fuel to the flame. Consequently, fuel diffusion ues of Ra∗ . It is seen that there exist minimum val-
plays a more important role than that of the oxidant, ues for Δ and Da∗ , Δc and Da∗E , and there are two
so that variation of LeF leads to a much larger shift branches of solutions when Δ > Δc and Da∗ > Da∗E ,
of xf . respectively. However, because of the nonlinear rela-
Fig. 3 shows variations of the coefficient of fuel tion between Δ and Da∗ , Δc does not coincide with
leakage in the total excess enthalpy, CF , with LeF Da∗E except for the case of zero excess enthalpy when
and LeO for different values of Ra∗ , respectively. LeF = LeO = 1 and Ra∗ = 0. It is noted that flame
Here only the results for CF are presented because extinction corresponds to the minimum of Da∗ , Da∗E ,
the flame is located on the oxidant side of the stag- which is a parameter that can be independently speci-
nation surface for the current problem, so that γ is fied. However, it does not correspond to the minimum
sufficiently larger than zero and hence SF  SO . It is of Δ, Δc , which is not an independently specifiable
seen that CF decreases with the increase of LeF , LeO parameter because it depends on the value of γ de-
H.Y. Wang et al. / Combustion and Flame 148 (2007) 100–116 109

termined from the leading order solutions. As shown


in Fig. 4b, the turning point Δc in the SF ∼ Δ plot,
indicated by the “×” symbol, is located on the up-
per branch of the SF ∼ Da∗ plot. The extinction limit
here, Da∗E , indicated by the “F” symbol, corresponds
to the kinetic extinction limit at high stretch rates.
Near this limit, increasing Da∗ leads to longer res-
idence time and hence more complete burning and
less reactant leakage. Thus, the lower branch is the
physically realistic solution. Fig. 4b also shows that
the SF ∼ Da∗ curve shifts to the right with increas-
ing Ra∗ . This means that for larger Ra∗ a higher Da∗
is required for steady burning in order to compensate
for the effect of radiative loss, and for a fixed Da∗ the
amount of fuel leakage SF with larger Ra∗ is greater Fig. 5. Variations of the extinction Damköhler number, Da∗E ,
than that with smaller Ra∗ . Consequently, the extinc- with the radiative loss, Ra∗ , for LeF = LeO = 1 and
tion limit, Da∗E , increases with Ra∗ . To illustrate the the Da∗ versus Ra∗ /Γ lines for different relative radiative
effect of neglecting the excess enthalpy on the pre- strengths, Γ .
diction of the extinction limit, the SF ∼ Da∗ curves
for different Ra∗ without considering the excess en- smaller Ra∗ leads to a smaller extinction Damköh-
thalpy are also plotted in Fig. 4b. It is observed that ler number, namely Da∗E,K , and hence corresponds to
the fuel leakage at the same Da∗ and the extinction the kinetic extinction limit, whereas the one at larger
Ra∗ corresponds to the radiative extinction limit oc-
Damköhler number, Da∗E , are underestimated due to
curring at a larger Da∗ , namely Da∗E,R . Thus, steady
the omission of the excess enthalpy, and this under-
burning is only possible within a limited range of
estimation becomes more significant with increasing
Damköhler number, Da∗E,K < Da∗ < Da∗E,R . Further-
Ra∗ . This is because the excess enthalpy is negative
more, it is seen that with increasing Γ the two ex-
for the parameters used in this figure (LeF = LeO = 1
tinction Damköhler numbers approach each other and
and Ra∗ > 0) and therefore tends to weaken the flame.
there is no intersection point once Γ exceeds the criti-
Thus omitting it would lead to flames that were more
cal value, Γc . This trend can be observed more clearly
resistant to extinction, with smaller values of Da∗E .
from the dual extinction Damköhler numbers versus
For a given Ra∗ , the extinction Damköhler num-
Γ plot shown in Fig. 6. It is seen that Da∗E,R de-
ber, Da∗E , can be solved from Eqs. (47) and (49).
However, because both Ra∗ and Da∗ depend on the creases and Da∗E,K increases with the increase of Γ ,
stretch rate k, the Da∗E solved for a fixed Ra∗ , such and the two branches of extinction Damköhler num-
as those in Fig. 4b, is only relevant for studying the bers ultimately merge at Γ = Γc so that burning is
effects of varying the fuel/oxidant system for a coun- not possible regardless of the stretch rate for Γ > Γc .
terflow of fixed stretch rate. On the other hand, an Furthermore, it is noted that the merging is mostly
equally relevant question to ask is the effect of varying effected by the reduction of Da∗E,R instead of the in-
the stretch rate for a fixed fuel/oxidant system. Be- crease of Da∗E,K . Thus, Da∗E,R is more sensitive to Γ
cause of the inverse dependence of both Ra∗ and Da∗ than Da∗E,K . This is because Da∗E,R is usually much
on k, the relative effect of radiative heat loss can be larger than Da∗E,K and hence increasing Γ implies a
evaluated by defining a parameter Γ = Ra∗ /Da∗ [6] much larger increase of radiative loss at the radiative
that represents the strength of the radiative loss rela- limit than at the kinetic limit. Consequently, the extent
tive to the chemical heat release. It depends only on of Da∗E,R decrease with increasing Γ is much larger
the thermochemical parameters, but is independent of than the extent of Da∗E,K increase. To illustrate the ef-
the stretch rate k. Thus, in additional to Eqs. (47) and fect of missing the excess enthalpy on the dual extinc-
(49), the extinction Damköhler numbers also need to tion limits, the result without considering the excess
satisfy Ra∗ /Da∗ = Γ at the same time and hence enthalpy is also shown in Fig. 6 by the dash curve. It
can be obtained from the intersection points between is seen that while neglecting the excess enthalpy only
the Da∗E ∼ Ra∗ curves and the Da∗ ∼ Ra∗ /Γ lines. has a very small effect on the kinetic extinction limit,
Fig. 5 shows the variation of Da∗E with Ra∗ for LeF = it has a much larger effect on the radiative limit. This
LeO = 1 and the Da∗ ∼ Ra∗ /Γ lines for different is due to the much larger radiative loss at the radiative
values of Γ . It is seen that for a fixed system with extinction limit. As shown in Fig. 3, the excess en-
Γ smaller than the critical value Γc ≈ 0.298, there thalpy is negative and decreases with the increase of
are two intersection points and hence two extinction radiative loss. Thus, the absolute value of the excess
limits at different level of radiative loss. The one at enthalpy at the radiative limit is much larger than that
110 H.Y. Wang et al. / Combustion and Flame 148 (2007) 100–116

Fig. 8. Variations of the normalized flame temperature


Fig. 6. Variations of the dual extinction Damköhler numbers
Tf /Tad with the Damköhler number Da∗ for LeF = LeO =
with the relative radiative strength Γ for LeF = LeO = 1
1 and different relative radiative strengths Γ .
with and without considering the excess enthalpy.

insufficient residence time is the dominant mecha-


nism for the excessive reactant leakage. Thus, increas-
ing Da∗ leads to a longer residence time and con-
sequently more complete burning and less reactant
leakage. However, for a fixed system, increasing Da∗
(decreasing the stretch rate) implies that the radiative
loss, Ra∗ , increases at the same time. Thus, at large
enough Da∗ , the radiative loss eventually becomes
dominant and is expected to lead to substantial re-
duction in the flame temperature, and hence increased
reactant leakage due to the substantial reduction in
the reaction rate. Thus it can be concluded that the
ultimate cause of extinction at the radiative limit is
still the insufficient time for the reaction to complete.
Moreover, it is seen that the fuel leakage SF at the
Fig. 7. Fuel leakage SF as a function of the Damköhler num- radiative extinction limit is smaller than that at the ki-
ber Da∗ for LeF = LeO = 1 and different relative radiative netic limit. This is because near the radiative limit,
strengths Γ . the flame is much weaker and hence is susceptible to
extinction with smaller reactant leakages. The upper
at the kinetic limit, and neglecting it as a factor weak- branch of the solutions, plotted in dotted lines, shows
ening the flame leads to an O(1) overestimation of the opposite dependence on Da∗ and thus is not phys-
Da∗E,R and a slight underestimation of Da∗E,K . For ex- ically realistic. The two branches of solutions form a
ample, at Γ = 0.26, missing the excess enthalpy leads closed isola-shape of SF ∼ Da∗ curve with two turn-
to a 60% overestimation of Da∗E,R and a 13% under- ing points instead of only one for the open C-shaped
estimation of Da∗E,K . curve shown in Fig. 4b.
Fig. 7 shows variations of the fuel leakage SF with Fig. 8 shows variations of the normalized flame
Da∗ for LeF = LeO = 1 and different values of Γ . It temperature Tf /Tad with Da∗ for LeF = LeO = 1 and
is seen that the solutions are bounded by the two turn- different values of Γ , where Tad is the adiabatic flame
ing points, with the one at smaller Da∗ , marked by temperature. It is seen that although the reactant leak-
“F,” representing the kinetic extinction limit, Da∗E,K , age decreases and then increases with Da∗ , as shown
and the one at larger Da∗ , marked by “!,” repre- in Fig. 7, the flame temperature decreases monoton-
senting the radiative limit, Da∗E,R . It is shown from ically because the flame suffers more radiative loss
the lower branch solution that with the increase of with larger Da∗ .
Da∗ from Da∗E,K , SF first decreases significantly, then We now study effects of the Lewis numbers of
levels off, and finally increases as Da∗ approaches the fuel and oxidant, LeF and LeO , on the extinction
Da∗E,R , leading to the second turning point at larger characteristics of the counterflow diffusion flame with
Da∗ . This trend is due to the fact that radiative loss radiative loss. Fig. 9 shows variations of the extinction
is rather small near the kinetic extinction limit and Damköhler number, Da∗E , with LeF and LeO for dif-
H.Y. Wang et al. / Combustion and Flame 148 (2007) 100–116 111

(a)
(a)

(b)

Fig. 9. Variations of the extinction Damköhler number Da∗E


(b)
with (a) LeF (LeO = 1) and (b) LeO (LeF = 1) for different
values of radiative loss, Ra∗ . Fig. 10. Variations of the ratio of the extinction Damköhler
numbers Da∗E /Da∗E,ad and the contributions from the effects
ferent values of Ra∗ . It is seen that for the adiabatic of radiative loss on the flame temperature, fT , the excess
flame (Ra∗ = 0), Da∗E increases monotonically with enthalpy, fh , and the reduced extinction Damköhler num-
increasing LeO and LeF , consistent with the results of ber, fΔ , with (a) LeF (LeO = 1) and (b) LeO (LeF = 1) for
radiative loss Ra∗ = 2 × 10−4 .
Seshadri and Trevino [19]. In the presence of radiative
loss, Da∗E increases monotonically with LeO , whereas
is seen from Eq. (50) that radiative loss modulates
it decreases and then increases with LeF . Specifically,
Da∗E through its effects on the flame temperature,
at small LeF , Da∗E is very sensitive to the radiative
loss, so that even a very small amount of loss, e.g., fT = (Tad /Tf )6 exp[Ta (1/Tf − 1/Tad )], excess en-
Ra∗ = 2 × 10−4 , is able to induce a very large relative thalpy, fh = exp(htotal,ad − htotal ), and the reduced
increase of Da∗E . However, at larger LeF this sensitiv- extinction Damköhler number, fΔ = Δext /Δext,ad .
ity becomes moderate, leading to the nonmonotonic Thus, it is necessary to evaluate their relative contri-
dependence of Da∗E on LeF for a fixed Ra∗ , as shown butions to Da∗E at different Lewis numbers and radia-
in Fig. 9a. The ratio of Da∗E to its adiabatic value, tive losses.
Da∗E,ad , can be obtained from Eq. (47) as Fig. 10 shows variations of Da∗E /Da∗E,ad , fT , fh ,

     and fΔ with LeF and LeO for Ra∗ = 2 × 10−4 . It


Da∗E Tad 6 1 1 is seen that the relative contributions from fT , fh ,
= exp Ta −
Da∗E,ad Tf Tf Tad and fΔ are different at small and large values of LeF
and LeO . Specifically, at small LeF , radiative loss in-
Δext
× exp(htotal,ad − htotal ) , (50) duces a very large increase of Δext , which contributes
Δext,ad mostly to the increase of Da∗E . This contribution de-
where the subscript “ad” designates the values cor- creases significantly with increasing LeF so that the
responding to the adiabatic state, i.e., Ra∗ = 0. It contribution from the temperature term, fT , becomes
112 H.Y. Wang et al. / Combustion and Flame 148 (2007) 100–116

dominant at large LeF . Over the entire range of LeF , extinction occurs. For example, for LeF = LeO = 1
the contribution from the excess enthalpy, fh , is sec- and Γ = 0.1, the kinetic and radiative extinction oc-
ondary. However, it contributes most to the increase cur at Ra∗ = 1.2 × 10−4 and 0.017, respectively. It
of Da∗E at small LeO , as shown in Fig. 10b. The con- is seen from Fig. 11 that the variations of fΔ and fh
tribution then decreases significantly with increasing with LeF and LeO exhibit the same trend as those for
LeO and becomes a secondary effect when LeO is suf- small radiative loss Ra∗ = 2 × 10−4 , whereas the rel-
ficiently large. The variation of fΔ with LeO shows ative contribution from the temperature term, fT , is
the opposite trend to fh . It is small at small LeO but much larger. This is due to the larger reduction of the
increases with increasing LeO and becomes compara- flame temperature under larger radiative loss. Further-
ble to fT at large LeO . Over the entire range of LeO , more, Figs. 10a and 11a show that it is the very large
fT plays a consistently important role, although it is increase of Δext from the adiabatic value at small LeF
a little smaller than fh at small LeO . The radiative that leads to a very large relative increase of Da∗E , as
loss, Ra∗ = 2 × 10−4 , used in Fig. 10 can be treated shown in Fig. 9a.
as a typical order of value at which kinetic extinction Because Eq. (36) shows that the minimum of Δ,
occurs. Fig. 11 shows variations of Da∗E /Da∗E,ad , fT , Δc , is determined by γ and in general the reduced
fh , and fΔ with LeF and LeO for Ra∗ = 0.01, which extinction Damköhler number, Δext , is close to Δc ,
can be considered as a typical value at which radiative it is instructive to explore the very high sensitivity of
Δext to radiative loss at small LeF by studying the
variations of γ with LeF . Fig. 12 shows variations
of γ and the corresponding Δext with LeF for dif-
ferent values of Ra∗ . It is seen that for the adiabatic
flame, γ is close to unity at small LeF , implying that

(a)

(a)

(b)

Fig. 11. Variations of the ratio of the extinction Damköhler


numbers Da∗E /Da∗E,ad and the contributions from the effects (b)
of radiative loss on the flame temperature, fT , the excess
enthalpy, fh , and the reduced extinction Damköhler num- Fig. 12. Variations of (a) γ and (b) the reduced extinction
ber, fΔ , with (a) LeF (LeO = 1) and (b) LeO (LeF = 1) for Damköhler number Δext with LeF for LeO = 1 and different
radiative loss Ra∗ = 0.01. values of radiative loss, Ra∗ .
H.Y. Wang et al. / Combustion and Flame 148 (2007) 100–116 113

(a)
Fig. 13. Variations of the extinction Damköhler number,
Da∗E , with radiative loss, Ra∗ , for LeF = 1, Γ = 0.1 and
different values of LeO and the Da∗ versus Ra∗ /Γ line with
relative radiative strength Γ = 0.25.

most of the chemical heat release is conducted to the


oxidant side and the flame is nearly adiabatic to the
fuel side. This is because at small LeF the flame lo-
cation is very close to the oxidant stream. It is seen
from Fig. 12b that in this case Δext is close to zero.
However, in the presence of radiative loss, which is
assumed to occur on both sides of the reaction zone
in the current study, the flame always has heat con-
ducted to the fuel side to compensate for the radiative
loss. Thus, γ decreases, leading to an increase of Δext
from its adiabatic value. Furthermore, because the (b)
adiabatic value of Δext is close to zero, even a small Fig. 14. Variations of the dual extinction Damköhler num-
increase of Δext from this value, induced by a very bers with relative radiative strength Γ for different values of
small amount of radiative loss, implies a very large (a) LeO (LeF = 1) and (b) LeF (LeO = 1).
relative increase. For example, at LeF = 0.5, with the
radiative loss Ra∗ = 2 × 10−4 , γ decreases from 0.98 LeO and LeF , respectively. It is seen that for the same
to 0.93, whereas Δext increases from 0.053 to 0.18, Γ the flammable range of Da∗ diminishes with the
which is about 3.4 times larger. increase of LeO , leading to a smaller Γc that the sys-
For a fixed system with radiative strength Γ , how- tem can sustain. Thus increasing LeO tends to weaken
ever, we have known that the extinction Damköhler the flame monotonically. Although the dependence of
number needs to satisfy Ra∗ /Da∗ = Γ at the same Da∗E on LeF is not monotonic over its entire range,
time. Thus, Fig. 9 alone is not enough to demonstrate Fig. 9a shows that, with an increase of LeF , Da∗E de-
variations of the dual extinction Damköhler numbers creases and increases monotonically for small and
with the Lewis numbers. In the same manner as that in large LeF , respectively. Thus, for the same reason,
Fig. 5, Fig. 13 shows variations of Da∗E with Ra∗ for it is seen from Fig. 14b that the flammable range of
LeF = 1, Γ = 0.1, and different values of LeO and the Da∗ diminishes as LeF increases from 1.5 to 2.5, and
Da∗ ∼ Ra∗ /Γ line with Γ = 0.25. It is seen that due extends as LeF increases from 0.6 to 0.8. It is noted
to the monotonic increase of Da∗E with LeO , increas- that Mills and Matalon [11] predicted monotonic de-
ing LeO leads to an increase of Da∗E,K and a decrease pendence of the flammable range of the Damköhler
of Da∗E,R , and hence a smaller flammable range of number on the Lewis number of the fuel, LeF . This is
Da∗ . Furthermore, it is expected that there are no in- because the radiative loss was assumed to be of O(ε)
tersection points between the Da∗E ∼ Ra∗ curve and quantity and hence does not affect the leading order
the Da∗ ∼ Ra∗ /Γ line if LeO is larger than a critical flame temperature. Thus radiative loss affects extinc-
value, indicating that steady burning is not possible if tion only through the excess enthalpy, which depends
LeO is too large. monotonically on LeF .
Fig. 14 shows variations of the dual extinction Fig. 15 shows variations of the dual extinction
Damköhler numbers with Γ for different values of Damköhler numbers with LeF for different values of
114 H.Y. Wang et al. / Combustion and Flame 148 (2007) 100–116

structure may not be directly affected by the partic-


ulars of the system. Thus, it is possible to perform
a generalized analysis of the reaction zone with the
influences of all the outside processes including ra-
diative loss coming in from the boundary conditions
through matching.
Second, this formulation is then applied to study
the extinction characteristics of the radiative counter-
flow diffusion flame with nonunity Lewis numbers of
the fuel and oxidant. In additional to the dual extinc-
tion limits, namely the kinetic limit at lower Damköh-
ler numbers and radiative limit at higher Damköhler
numbers, identified in previous analyses [5–9,11], the
current study has also gained some additional under-
standing, especially of the effects of nonunity Lewis
Fig. 15. Variations of the dual extinction Damköhler num- numbers. It is found that the kinetic extinction limit
bers with LeF for different values of LeO and relative radia- is minimally affected by radiative loss so that extinc-
tive strength Γ = 0.1. tion occurs close to the adiabatic flame temperature,
while a substantial amount of heat loss is associated
LeO and Γ = 0.1. It is seen that the flammable range with the radiative limit, so that the flame temperature
of Da∗ decreases with increasing LeO and LeF that at extinction is significantly reduced from the adia-
sufficiently large. When they are both large enough, batic value. Reactant leakage, however, is ultimately
steady burning is not possible for this value of Γ . the root cause for both limits, with the extent of leak-
With the increase of LeO , the maximum value of LeF age being greater for the kinetic limit. The flammable
the system can sustain decreases. Fig. 15 indicates range of the Damköhler number decreases monoton-
that steady burning for radiative diffusion flames is ically with increasing LeO , indicating that increasing
only possible within a limited range of Lewis num- LeO tends to weaken the flame. However, this range
bers. This is because flame extinction is purely ther- shows nonmonotonic dependence on LeF in that it in-
mal in nature under the assumption of a one-step re- creases and decreases with increasing LeF when it is
action and since thermal conduction and radiation are sufficiently small and large, respectively. This non-
both heat loss mechanisms for flame extinction; either monotonic dependence is due to the very high sensi-
of them being large enough is able to induce extinc- tivity of the extinction Damköhler number to radiative
tion. loss at small LeF . The root cause for this phenomenon
is that the flame is located very close to the oxidant
stream at small LeF so that most of the chemical heat
4. Conclusions release is conducted to the oxidant side and the flame
is nearly adiabatic on the fuel side. A small amount of
The present study has yielded the following spe- radiative loss can lead the flame to deviate from this
cific contributions and understandings regarding the condition and hence can lead to a rather large relative
subject phenomena. First, we have performed a mul- increase of the extinction Damköhler number. The in-
tiscale asymptotic analysis for counterflow diffusion fluence of radiative loss on the extinction Damköh-
flames with flame radiative heat loss and developed a ler numbers is found to be through its effects on the
formulation for general Lewis numbers of the fuel and flame temperature, excess enthalpy, and the reduced
oxidant with the proper consideration of the excess extinction Damköhler number, and their relative con-
enthalpy that was overlooked in previous analyses. tributions are different under different radiative loss
This formulation is expressed by the jump relations and Lewis numbers of the fuel and oxidant. In most
in terms of the outer solutions and the reactant leak- cases, the contribution from the flame temperature is
ages through the reaction zone. The reactant leakages the largest and its relative importance increases with
are obtained from solving the canonical form of the increasing radiative loss. The contributions from the
structure equation and hence can take advantage of other two, however, are also important in some cases.
the previous results, such as those of Cheatham and At small LeF , when the flame is located very close
Matalon [17]. The reason that the structure equation to the oxidant stream, the contribution from the re-
of the reaction zone, with a radiation zone sandwich- duced extinction Damköhler number is large and even
ing it in the current analysis, still can degenerate to dominant in the case of small radiative loss. The con-
the Liñán canonical form [1] is that the thin reaction tribution from the excess enthalpy is important for
zone is reactive–diffusive in nature. Consequently, its small LeO and it may be comparable to that from the
H.Y. Wang et al. / Combustion and Flame 148 (2007) 100–116 115

flame temperature when radiative loss is small. Thus, crucial element in the stability analyses. For exam-
overlooking the excess enthalpy in previous analyses ple, Kim et al. [18] and Cheatham and Matalon [17]
may result in rather large error in the predicted extinc- and Kukuck and Matalon [14] respectively show that
tion Damköhler numbers, especially the kinetic one. flames exhibit cellular and pulsating instabilities for
Moreover, it is found that for a fixed relative radiative positive and negative values of the excess enthalpy.
strength, steady burning may not be possible when the
Lewis numbers of the fuel and oxidant are too large.
Finally, it should be noted that the above discussions Acknowledgment
are meant for large initial mixture strengths such that
the flame is located in the oxidant side of the stag- This work was supported by NASA and the Air
nation surface. The roles of LeF and LeO should be Force Office of Scientific Research, under the techni-
interchanged for small initial mixture strengths under cal monitoring of Dr. Kurt Sacksteder and Dr. Julian
which the flame is located in the fuel side of the stag- Tishkoff, respectively.
nation surface.
Third, the relations (37)–(41), developed in this
study for the counterflow diffusion flame with flame Appendix A
radiation as a demonstration problem, can be applied
to any one-dimensional configuration with Eq. (41) Integrating Eq. (23) and applying the matching
revised as conditions, Θ2± |ζ →±∞ ∼ −∞ and Θ2± (0) ∼ 0, yield
  
 ±   ±
Buf T1 −
1 ∂T1
= − uf YF,1 − LeF −1 ∂YF,1 dΘ2± ∂T0 (xf ) 2 Θ2
A ∂x ∂x =∓ + 2RaTf exp
dζ ∂x Tf

∂Y (A.1)
= − uf YO,1 − Le−1
O,1
O ,
∂x and
where uf is the flow speed at the flame sheet. dΘ2± (0) ∂T0± (xf )
Furthermore, although this formulation is applied = A± . (A.2)
dζ ∂x
exclusively to the extinction analysis in the current
study, it is also applicable to the analyses of thermal– Multiplying Eq. (24) by dΘ2± /dζ , integrating once,
diffusive instabilities in a manner similar to those in and applying the matching conditions (12) for dΘ1± /
dζ , dΘ2± /dζ , and Θ2± as ζ → ±∞ yield
Cheatham and Matalon [17] and Kukuck and Mat-
alon [14]. For example, we introduce small perturba-
 ±
tions in the form dΘ1± (0) dΘ2± (0) Θ2
    = RaΘ1± exp . (A.3)
(T , YF , YO ) = T̄ , ȲF , ȲO + εα u(x), v(x), w(x) dζ dζ Tf

× exp(ik1 y + σ t), (51) Substituting Eq. (23) into (A.3) to replace the term
Ra exp(Θ2± /Tf ) with d 2 Θ2± /dζ 2 , integrating once,
where the overbar “−” designates steady state solu- and applying the matching condition (12) for Θ1± and
tions, y is the transverse direction, k1 the wavenum-
Θ2± as ζ → ±∞ yield
ber in this direction, σ a complex number whose real
 
part identifies the growth rate of the perturbation, α
± T1± (xf ) dΘ2±
a small parameter, and u, v, and w the perturbations Θ1 =
for T , YF , and YO , respectively. Substituting Eq. (51) ∂T0± (xf )/∂x dζ

into the governing equations (1)–(3) and Eqs. (37)–   ±
 2RaTf Θ2
(41) yields the governing equations and the jump and = T1± (xf ) 1 + ± exp
(∂T0 (xf )/∂x) 2 Tf
leakage conditions for the perturbations, from which
the dispersion relation relating the growth rate to the (A.4)
parameters describing the combustion system, such as and
the Damköhler number and the Lewis numbers of the
fuel and oxidant, is formed. Because this model is Θ1± (0) = A± T1± (xf ). (A.5)
able to predict the dual extinction limits of radiative Integrating Eq. (25) once in the same manner yields
diffusion flames, it is expected to predict the neutral
!
stability boundaries near both the kinetic and radiative dΘ2± dΘ3± dΘ1± dΘ2±
extinction limits. = −xf dζ
dζ dζ dζ dζ
Finally, it should be emphasized that although the  ±
Θ2
inclusion of excess enthalpy does not lead to qual- + RaΘ3± exp + C±, (A.6)
itative differences in the extinction results, it is a Tf
116 H.Y. Wang et al. / Combustion and Flame 148 (2007) 100–116

where C ± are integral constants. Substituting [2] U. Bonne, Combust. Flame 16 (1971) 147–159.
Eq. (A.4) into (A.6) to replace the term dΘ2± /dζ by [3] M. Sibulkin, A.K. Kulkarni, K. Annamalai, Proc. Com-
 ∂T0± (xf )/∂x  ± bust. Inst. 18 (1981) 611–617.
± Θ1 and applying the matching condi- [4] S.H. Sohrab, A. Liñán, F.A. Williams, Combust. Sci.
T1 (xf )
tion (12) for Θ1± , Θ2± , and dΘ3± /dζ as ζ → ±∞
Technol. 27 (1982) 143–154.
[5] J.S. T’ien, Combust. Flame 65 (1986) 31–34.
lead to
[6] B.H. Chao, C.K. Law, J.S. T’ien, Proc. Combust. Inst.
 
∂T ± (xf ) ∂T1± (xf ) 1 23 (1990) 523–531.
C± = 0 + xf T1± (xf ) [7] T.K. Oh, J.S. Lee, S.H. Chung, Int. J. Heat Mass Trans-
∂x ∂x 2
fer 37 (1994) 2893–2900.
and [8] F. Liu, G.J. Smallwood, O.L. Gulder, Y. Ju, Combust.

 Flame 121 (2000) 275–287.
dΘ3± 1 ±
 2RaTf exp(Θ2± /Tf ) [9] K. Maruta, M. Yoshida, H. Guo, Y. Ju, T. Niioka, Com-
= − xf T1 (xf ) 1 +
dζ 2 (∂T0± (xf )/∂x)2 bust. Flame 112 (1998) 181–187.
[10] J.S. Kim, F.A. Williams, J. Eng. Math. 31 (1997) 101–
 ±
T1± (xf ) ± Θ2 118.
+ ± RaΘ 3 exp [11] K. Mills, M. Matalon, Proc. Combust. Inst. 27 (1998)
(∂T (xf )/∂x) 2 Tf
0 2535–2541.
∂T1± (xf )/∂x + 12 xf T1± (xf ) [12] R.H. Chen, G.B. Mitchell, P.D. Ronney, Proc. Com-
+ " . bust. Inst. 24 (1992) 213–221.
2RaTf exp(Θ2± /Tf )
1+ ± 2
[13] M. Füri, P. Papas, P.A. Monkewitz, Proc. Combust. Inst.
(∂T0 (xf )/∂x) 28 (2000) 831–838.
Then, applying the matching conditions, Θ2± (0) ∼ 0 [14] S. Kukuck, M. Matalon, Combust. Theory Model. 5
(2001) 217–240.
and Θ3± (0) ∼ 0, yields [15] C.H. Sohn, J.S. Kim, S.H. Chung, K. Maruta, Combust.
  Flame 123 (2000) 95–106.
dΘ3± (0) 1 1
±
1 ∂T (xf )
∼ A± − ± T1± (xf ) + ± 1 . [16] M. Miklavčič, A.B. Moore, I.S. Wichman, Combust.
dζ 2 A A ∂x Theory Model. 9 (2005) 403–416.
(A.7) [17] S. Cheatham, M. Matalon, J. Fluid Mech. 414 (2000)
105–144.
[18] J.S. Kim, F.A. Williams, P.D. Ronney, J. Fluid Mech.
References 327 (1996) 273–301.
[19] K. Seshadri, C. Trevino, Combust. Sci. Technol. 64
[1] A. Liñán, Acta Astronaut. 1 (1974) 1007–1039. (1989) 243–261.

You might also like