Chapter 5 Nonpremixed Microflame Edited
Chapter 5 Nonpremixed Microflame Edited
The advancement of human civilization is closely related to the ability to use and
control flames. Traditionally, flames have been used for heating, cooking,
illumination, and power generation for industrial and civil applications.. Modern
applications of flames include power generation for electricity and transportation, and
Recently, with increasing demands of micro devices such as micro satellites and
micro aerial vehicles, needs for a micro power source to activate these systems have
significantly increased [1]. These micro systems require high-density power sources
for long periods of operation. The energy density of typical hydrocarbon fuels is about
100 times higher than that of the most advanced batteries. Even with heat losses in the
process of extracting power from burning the fuel, a micro-scale combustion system
has been considered as a viable alternative to batteries. To develop such combustion
importance.
For micro-scale applications, premixing fuel and air is not favored as the process
microjet diffusion flames. Numerical and advanced non-intrusive diagnostic tools are
diffusion flames and microjet diffusion flames are is given in the following sections.
devices. The ability to predict the coupled effects of complex transport phenomena
with detailed chemical kinetics in these systems is critical for modeling reacting
flows, improving engine efficiency, and understanding the phenomena such as flame
extinction and pollutant formation. Laminar jet diffusion flames are fundamental to
combustion and have received much attention as a model system due to their
relatively simple geometry and fluid mechanical characteristics. Efforts have been
Laminar jet diffusion flames fall into three categories: (1) the Burke--–Schumann
flame [13] controlled by diffusion, (2) the Roper flame [14] controlled by momentum
or buoyancy, and (3) the microflame [15] controlled by diffusion or momentum. The
first two types of flames have long been extensively investigated [16]. The third type
of flame was only recently investigated [15, 17-–23]. However, these studies were
concerned with bulk flame geometrical and thermal characteristics. Due to its small
size and negligible buoyancy force, the microjet flame can be used as a point heat
source with no preference to the orientation in future micro power devices. Therefore,
it is necessary to determine the smallest stable flame that can be maintained and to
numerically studied [17]. The importance of axial diffusion in the theoretical model
and the insignificance of buoyancy effect on the height and shape of the microjet
diffusion flame hashave been identified [15]. More recent investigations of microjet
diffusion flames include those of various fuels: methane [19-–, 20, 23-–, 24], propane
[21, 22], and hydrogen [25] flames. These studies found that the microjet flame hadis
microgravity flames [26]. Also, due to its small size, the heat loss to the burner is
relatively large. Thus, the flame might always operate in a severe, near-extinction
condition. It is also known that different fuels produce different visible flame shapes
near the burner port. For example, the C 2 class of hydrocarbon [15] and hydrogen [25]
flames extend upstream of the burner port, while the propane [21, 22] and methane
yield stand-off flames [19-–, 20, 23-–, 24]. Experimental results of flame shapes of
microjet methane diffusion flames are different using the direct photograph and laser
shadowgraph [20]. This implies that an intensively hot region may exist just ahead of
vertical straight stainless-steel tubes with the inner diameter (d) ranging from 186 to
778 µm. The tube wall thickness is 79, 88, 122, 147, 148, and 124 µm for with d =
186, 212, 324, 382, 529, and 778 µm, respectively. Methane fuel is introduced
through the tube into the quiescent atmospheric air. The flame shapes are recorded
using a color CCD (Charge-coupled device) camera with a macro lens. Figures 5.1
predicted results clearly indicate the characteristic flame structure of the microjet
diffusion flame, which severely quenches locally at the flame base and the flame
5.32. Methodology
5.32.1 Experimental Setup
Because of the small flame sizes, intrusive techniques are unsuitable for the study of
microjet diffusion flames. Non-intrusive optical methods are used for the present
The laminar microjet diffusion flames investigated here are stabilized on a vertical
straight stainless-steel (AISI 304) tubes. The fuel flow is fully developed when it exits
the burner. The fuel flow rate is measured with an electrical mass flowmeter
(BROOKS Brooks 5850E), which is well calibrated using a soap-bubble flow meter.
processing. Micro flame images are obtained by a highly sensitive 3-chip color CCD
camera (Sony DXC-9000) with external triggering and recorded in the NTSC format.
The camera shutter time can be adjusted for best images of various flame conditions.
The captured images are then digitized for further analysis. The visible flame heights
have previously been measured using an identical imaging system by Cheng et al.
CCD camera (Cooke SensiCam) with a macro lens and a narrowband interference
For microjet diffusion flames, detailed measurements are difficult to achieve because
of the tiny flame sizes. The sizes of microjet flame in this study are only several
millimeters and the flame thicknesses are on the order of micrometers. Neither
intrusive nor non-intrusive measurements can obtain sufficient resolution for species
simulations [27] with sufficiently small grid resolution are used to comprehend the
flame structures and characteristics. Theoretical models are also applied to examine
To numerically model the laminar microjet diffusion flames, the governing equations
of mass, momentum, energy, and chemical species for a steady axisymmetric reacting
(5.2)
(5.3)
(5.4)
where ρ, p, T, Y, cp, h, w, R0, M, gx, and v = (u, ) are the density, pressure,
temperature, mass fraction, specific heat capacity of the mixture, enthalpy, species
the x-direction, and velocity vector, respectively. µ, λ, and D are the viscosity, thermal
(5.5) stands for the i-th chemical species. The second term in the bracket of Equations
(5.3) and (5.4) is the thermo-diffusion or Soret diffusion due to the effect of
(5.6)
where Dij is the binary diffusion coefficient. The binary mass diffusivity is determined
(5.7)
where M is the mixture molecular weight and kij is the thermo-diffusion ratio.
conditions is shown in Figure. 5.3. The governing equations are solved using a
for solving the discretized equations with a control volume formulation in accordance
Consistent) [28] algorithm. The momentum equations are solved using the second-
order upwind scheme while the central difference method is used for the energy and
species equations. The above equations are solved along the mesh lines in the
transport data is obtained from the CHEMKIN package [289] and then the code
calculates the thermal conductivity and viscosity of the mixture using Wilke’s
formula. Thermal diffusion and buoyancy are included in this analysis but radiation
heat loss is neglected. A skeletal chemical kinetic mechanism [2930] is coupled with
the CFD package computational fluid dynamic program to predict the shape, length,
and structure of the micro methane flame. This skeletal chemical mechanism consists
sufficiently accurate for predicting flame speeds, extinction limits, and the thermo-
GRI-Mech 3.0 [301], is used to investigate the detailed flame structure and
stabilization mechanism.
It has been shown [15, 21, 312] that simple models derived from similarity analysis
can adequately predict flame height and flame shape for laminar jet diffusion flames.
diffusion flames. For a steady, axisymmetric, vertical laminar jet at low Mach
numbers, uniform pressure, negligible buoyancy, and negligible mass diffusion, heat
conduction as well as viscous action in the axial direction but fast chemical reaction
rates, the governing equations (5.1)-()–(5.4) can be simplified and represented as the
The dimensionless axial velocity component (u*) and mixture fraction (f) become
(5.9)
The flame shape can be determined from Equation (5.9) by assuming that the flame
(5.10)
(5.11)
where is the dimensionless flame height, fst is the stoichiometric value of the
mixture fraction (0.055 for CH4--air), Re is the Reynolds number based on the jet
diameter, , and .
Equations (5.10) and (5.11) are derived by assuming negligible axial diffusion in the
flame. The Fflame shapes calculated including the axial diffusion term waswere
shown to better agree with thosethat measured in experiment for low Reynolds
number C2 class hydrocarbon microflames [15]. If the axial diffusion term is taken
(5.12)
with the far-field boundary conditions F = 0 and F′= 0 as ζ→ ∞. The flame shape is
Equation (5.10).
In the study of microjet propane flames, Matta et al. [21] postulated that flame
behaviors are primarily controlled by diffusion and that buoyancy becomes less
important, at least foer cases near the quenching limit. Therefore, they adopted the jet
diffusion flame model of Turns [323] that does not include buoyancy effects for the
flame length predictions. The flame length depends only upon the volumetric fuel
(5.13)
where D is the mass diffusivity and YF,stoic is the stoichiometric fuel mass fraction. In
addition, Matta et al. [21] hypothesized that the critical fuel flow rate at quenching is
that at which the predicted flame length equals the measured standoff distance.
Equating the flame length in Equation (5.13) to the measured standoff distance yields
For the estimation of jet diffusion flame lengths, Roper [14] modified the Burke--
Schumann theory to allow the mass velocity of fuel gas to vary with axial distance as
affected by buoyancy and in accordance with continuity. For the circular burner port,
(5.14)
coefficient evaluated for the oxidizer at the oxidizer stream temperature, TO, and Tf is
the mean flame temperature. Equation (5.14) can be applied to estimate the flame
length regardless of whether or not buoyancy is important, and is applicable for fuel
jets emerging into either a quiescent oxidizer or a co-flowing stream, as long as the
The solutions of jet diffusion flame theory for the flame length estimations are mainly
dependent upon the volumetric flow rate and hence on the Reynolds number. Chung
and Law [334] extended the Burke--Schumann theory to include the effects of both
streamwise (axial) and preferential diffusion and derived a solution for the flame
length prediction. The solution that is independent of the Reynolds number but
(5.15)
where c is the normalized half-width of the inner wall, YO0 is the mass fraction of the
oxidizer at the oxidizer stream, αn = (Pe2 + 4π2n2)1/2, and Zf* is the normalized flame
height. Equation (5.15) suggests that with the limit of Pe → 0 the flame becomes
significant than streamwise convection. This, however, has not been experimentally
validated.
Equations (5.10), (5.13), (5.14), and (5.15) are all derived by assuming unity Lewis
(Le) and Schmidt (Sc) numbers. If Sc is not equal to unity, 0.704 for methane, then the
, (5.16)
where x* is the flame shape normalized by the nozzle diameter and the flame length is
In the present study, Equations (5.8) (without axial diffusion) and (5.12) (with axial
diffusion) as well as Equation (5.16) are solved numerically to predict the flame
shapes and flame lengths. Equations (5.13), (5.14), and (5.15) are used only for the
and theoretical results are made to assess the applicability of the simple jet flame
are solved numerically to obtain the flame shape. The calculated flame shapes with
the measured and predicted results of Ban et al. [15] for the d = 186 µm flames
operated at several fuel flow rates are compared in Figure. 5.4. The measured flame
shapes are indicated by symbols, the solid and dashed lines denote those from the
calculation with and without axial diffusion, respectively, and the dashed-dot-dashed
lines depict the calculated results of Equation (5.16). Results of Figure. 5.4 indicate
that flame shapes calculated with and without axial diffusion are in poor agreement
with the experimental data and the flame heights are also significantly over-predicted
by nearly an order of magnitude for all ranges of fuel flow rates. With the
heights are slightly improved. The comparison shown in Figure. 5.4 clearly
demonstrates that the simple jet flame theory fails completely in the prediction of the
flame shapes of microjet methane flames. For the d = 324 µm flames (not shown
here), similar discrepancies between the predicted and measured flame shapes are also
observed. The failure of the simple jet flame model may be attributed to the failure of
the similarity assumption, which is used in the derivation of the simple model, as
Equations (5.11) and (5.16) indicate that the flame length depends only upon the
Reynolds number. It means that different diameter tubes with the same fuel exit
Reynolds number will produce the same flame length. Although the Reynolds number
isdoes not explicitly appeared in Equations (5.13) and (5.14), the flame length can still
be related to the Reynolds number through the dependence on the flow rate (Qf).
Figure. 5.5. Figure 5.5(a) shows that the measured nondimensional flame length data
collapse into a straight line and can be scaled with the Reynolds number. Comparison
of the measured and calculated data indicates that the models of Ban et al. [15] , Turns
[323] , and Lee and Chung [345] all result in over-prediction of the flame lengths.
Figure 5.5(a) indicates that the flame length predicted by Roper’s model [14] agrees
very well with the measured data by assuming a mean flame temperature of 1500 K in
Equation (5.14). The slightly difference between the predictions and measurements
number. It should be noted that the model of Chung and Law [334] was derived for a
is slightly different from that used in the present study ( ), where b is the
half-width of the outer wall and lD is the radial position of the flame edge. The
Both measurement and theoretical data show that in the limit of Pe → 0 the flame
length becomes independent of Pe. This suggests that the streamwise diffusion can be
From these comparisons, one may expect that buoyancy acceleration would be more
(5.17)
where D is the fuel gas mass diffusivity estimated at the mean flame temperature and
a ≅ 0.6g[(Tf / T0)] is the mean buoyant acceleration [14]. If NDB << 1 then the flame is
buoyancy controlled. If NDB >> 1 then the flame is diffusion controlled. Similarly, NDM
(5.17)
where , uzo uz0 = Qf I YF,stoic / A is the ratio of the actual initial momentum flux to that
for uniform flow (for parabolic exit velocity I = 1.5 and for uniform flow I = 1.0) and
A is the cross-section area of the burner port. If NDM << 1, then the flame is
momentum controlled. If NDM >> 1, then the flame is diffusion controlled. It is noted
that the Froude number, based on the flame length instead of burner port diameter as
number, the Froude number, the diffusion-to-buoyancy parameter, and the diffusion-
flames. The calculated parameters, based upon experimental data and baseline data (Tf
= 1500 K, T0 = 300 K, I = 1.5, YF,stoic = 0.055, and D = 2.8195 cm2/s) are listed in
Table 5.1 for the d = 186 and 324 µm flames. It is clear that in the analysis
considering only diffusion and buoyancy, the flames are more in the diffusion-
dominated regime for Qf ≤ 5 cc/min because NDB is greater than 1. For Qf ≤ 5 cc/min
the microjet flames are in spherical shape. When Qf > 6 cc/min, the buoyancy
acceleration would be more important than molecular diffusion in the microjet flames
because of NDB < 1. The Froude number for the flames studied here ranges from
O(10-–1) to O(1). This indicates that the flames fall within the transitional to
velocity only for Qf ≤ 5 cc/min. This is in consistent with the NDM parameter analysis.
From the above parameter analysis, we can conclude that the microjet diffusion
flames studied here are not completely buoyancy-free and the molecular- diffusion is
To investigate the effect of tube size on extinction behavior, different tube diameters
are used. Figure 5.6 shows photographs of flames operated at fuel exit velocity near
extinction for tube diameters varying from 186 to 778 µm. The most notable feature
of Figure. 5.6 is that the standoff distance is essentially the same, approximately 0.78
mm, for all the tubes. In addition, the flame shapes are remarkably similar over the
range of tube diameters. It has been demonstrated in the previous section and shown
in Figure. 5.5 that the measured and predicted flame length is a linear function of
Reynolds number. Accordingly, a flame can not be sustained if the predicted flame
length for a fuel exit velocity is smaller than the measured standoff distance. Thus, it
is hypothesized that the critical fuel exit velocity at quenching is that at which the
predicted flame length equals the measured standoff distance [21]. Since Roper’s
model [14] is shown above to well predict the flame lengths, the model is further used
for quenching velocity predictions. The quenching velocity for different tube sizes is
determined by equating the flame length in Equation (5.14) to the measured standoff
in Figure. 5.7. The predicted quenching velocities (dashed line) are in excellent
agreement with the measured data. The good agreement between the calculated and
measured quenching velocities is due that near the quenching limit,; the microjet
flames are mainly diffusion- controlled and buoyancy has a minor effect on the flame
length. It is also found that the measured quenching velocities follow Re × d =
proposed [20] but different from Re × d2 = C that was obtained numerically for the
In order to clarify the effect of tube materials on the standoff distance, the computed
OH isopleths with different wall thermal conductivity are compared in Figure. 5.8 for
flames of d = 186, 324, and 529 µm. The standoff distance is measured along the jet
centerline from the burner exit to the bottom of blue flame cap. For the same tube
diameter, for example d = 186 µm, the image of Figures. 5.8(a), (d), and (g) are
(quartz tube), and k =1 W/m.K, respectively. It can be seen from Figures. 5.8(a)-–(c)
that the computed standoff distance (0.8 mm) is in good agreement with the measured
data (0.78 mm) for the stainless- steel tube. The maximum variation of the calculated
standoff distance is less than 3% for the same tube material with different tube
diameters and that is less than 5% for a fixed tube diameter with different tube
materials. Therefore, it can be concluded that the change of tube materials only has a
negligible effect on the predicted standoff distance, but does influence the quenching
gap between the flame and the tube. It is noted that the detachment of the flame base
from the burner rim is reduced as the thermal conductivity of the tube is reduced. This
is due that reduced thermal conductivity results in less conduction of heat from the
flame edge to the tube, and hence reduces the quenching gap. If an adiabatic wall
condition is imposed, the flame would attach to the burner and no quenching gap is
produced.
In order to have a further look into the detailed flame structure of the microjet
velocity of 2.39 m/s is adopted. This flow rate is just slightly above
Figure. 5.9. Note that the CH* image is Abel inverted and it is performed only for the
flame height and flame shape measurements. The CH* intensity can not be correlated
to the CH concentration because the CH* reaction mechanisms are not included in the
very good agreement. Both experimental and numerical results indicate that the
flame is quenched by the tube wall and creates a gap to allow oxidizer entrainment.
As a result, the flame stands off from the tube. The characteristics of semi-spherical
flame shape, non-buoyancy flame configuration, and large quenching gap between the
flame base and the tube are very similar to those of a microgravity flame, but the heat-
release rate is about two orders of magnitude larger than that of a microgravity flame
The computed results of the 2-D temperature, CH, O 2, CH4, CO2, H2O, H2, and CO
species mass fraction contours as compared with mixture fraction contours of lean (ξ
= 0.029), stoichiometric (ξ = 0.055), and rich (ξ = 0.089) limits of the methane flame
(5.18)
where YC, YH, and YO are the mass fraction for carbon, hydrogen, and oxygen atoms,
Wi is the atomic weight of species i, and subscripts f and O refer to fuel jet and
ambient air, respectively. It can be seen that a small amount of O 2 is entrained into the
standoff region from the gap between the burner wall and flame base and a small
amount of CH4 has diffused upstream of the burner port. The entrainment could
results in partial premixing of fuel and oxygen over the standoff distance. The
computed temperature contour shows that the maximum flame temperature locates at
the jet centerline near the stoichiometric mixture fraction contour and the maximum
of the CH isopleths. The unburned mixtures, the burner wall, and the fuel stream are
heated to a temperature higher than 700 K. This suggests that the standoff and the
flame stabilization are strongly related to the characteristic hot zone and heating of the
fuel stream and unburned mixture through the tube wall. The appearance of more
stable species of CO and H2 and products of CO2 and H2O within the standoff region
may further suggest that the high temperature in the standoff region is due to heat-
flame. However, the axial distributions of reaction rates (not shown here) and species
mass fractions along the jet centerline indicate that the final-product formation
reactions occur in the downstream region of the maximum temperature location and
no double flame structure exists in this region. Therefore, this high temperature in the
standoff region is most likely due to molecular heat conduction from the flame or
The computed axial profiles of temperature and selected species along the centerline
of the jet are shown in Figure. 5.11. The vertical dashed line denotes the location of
stoichiometric mixture fraction. It can be seen that oxygen is entrained and diffused
into the jet center near the burner exit and its concentration decreases with decreasing
located. This is usually seen in a normal laminar diffusion flame except that no O 2 is
present upstream of the intersection point in a diffusion flame. The distribution of CO,
H2, H2O, and CO2 is different from a typical diffusion flame structure [323] ,
only one CH peak occurring at the maximum temperature position (Figure. 5.11(b)).
This is also consistent with the computed pure CH4 diffusion flame structure [3899] .
To further examine the role of a tribrachial flame in the microjet flame stabilization,
the computed velocity vectors coupled with CH and stoichiometric mixture fraction
contours are depicted in Figure. 5.12 for a small region near the burner exit (0.8 mm ×
1.2 mm). Note that the computational domain is 4 mm × 16 mm in the radial and axial
direction, respectively. The maximum fuel velocity is 9.7 m/s at the centerline of the
burner exit. The velocity vectors show lateral expansion and longitudinal acceleration
as the fuel emerges from the burner exit and approaches the hot zone around the flame
base due to thermal expansion. The flow velocity in the ambient air as well as in the
downstream of flame base remains relatively constant and parallels to the axial
direction, indicating a negligible buoyancy effect. Also, a small amount of air is
entrained into the standoff region from the gap between the burner wall and the flame
base. Figure 5.12 indicates that the flow velocity (0.2 m/s) near the flame edge is less
than the laminar burning velocity (0.4 m/s) and no increase of velocity is observed
after passing the flame edge similar to the observation by Puri et al. [394040]. (2001).
In addition, the flame edge is not located on the stoichiometric mixture fraction point
[4011] . All of these outstanding features suggest that a tribrachial flame may not
exist in a microjet flame near extinction and the flame stabilization must be due to
From the above discussion one might conjecture that the stabilization of a standoff
under adiabatic and variable wall temperature conditions. With the adiabatic wall, the
flame zone extends upstream of the port and connects to the tube as indicated by CH
contours, while with the variable wall temperature condition (a realistic situation), the
flame is quenched on the tube wall and creates a gap to allow oxidizer entrainment.
As the flame is quenched by the tube wall, heat is transferred through the wall to
radicals in the vicinity of the exit. As usual, the HO 2 radicals are brisk in the chain-
terminating reactions when the flame is quenched on the wall. Detailed examination
of the intermediates indicates that the HO2 radicals play an important role in
HO2 radicals in contrast to the stoichiometric mixture fraction location. The CH mass
fraction isopleths and mass flux vectors of O and H radicals are shown on the left
hand side of the figure, while the total heat-release rate isopleths, HO 2 contours, and
mass flux vectors of OH and HO2 on the right hand side. It can be seen that the
maximum HO2 appears in the gap region then decreases to connect with heat-release
rate isopleths. The heat-release rate isopleth depicts a peak reactivity spot (called
reaction kernel [3566] ) located on the lean side of the stoichiometric mixture fraction
with the position of x 0.33 mm and r 0.6 mm. Hydrogen and oxygen atoms as
well as hydroxyl radicals diffuse to both sides of the flame zone and upstream against
the incoming fuel-rich flow, while the HO 2 radicals predominantly diffuse and
displace in the inward direction to the fuel stream near the tube exit.
To further delineate the stabilization process of the microjet flame, the detailed flame
structures across the standoff region (x = -–0.1 mm) and the reaction kernel (x = 0.33
mm) are examined. The radial profiles of the temperature, species mass fractions,
formation rates, net reaction rates, and net heat-release rates of the major elementary
steps at these two characteristic heights are illustrated in Figures. 5.15 and 5.16. In the
figures, the radial location of the tube wall and the stoichiometric mixture fraction (ξ
= 0.055) are indicated by the solid and dashed lines, respectively. For the case in
Figure. 5.15 (x = -–0.1 mm), the distributions of temperature and major species mass
fractions are typical of diffusion flames, except for the appreciable amount of
penetration of O2 to the fuel-side and the efflux of CH4 to the air-side through the gap
(Figure. 5.15(a)). The peak temperature and the production rates of minor species are
consumption rate and the production rates of H 2O and CO2 are also on the air-side
(Figure. 5.15(b)). These facts suggest that in the quenching gap region lean chemical
reactions take place. Figures 5.15(c) and 5.15(d) show the net reaction and heat-
release rates of major elementary steps. It is obvious that the most significant reaction
↔ H + H2O (R84) reactions to form the product H 2O and to further build up the
radical pool. Although the rate for H production through OH + H 2 ↔ H + H2O (R84),
significant than that for OH formation through HO 2 + CH3 ↔ OH + CH3O (R119) and
relatively low temperature region. In addition, the formation (R168) and destruction
(R119) of HO2 reactions are believed to play an important role for the hot zone and
reaction kernel formation in the quenching gap region. Figure 5.15(d) indicates that
the major contributors to the positive heat release are the reactions, including O + CH 3
OH + HO2 ↔ O2 + H2O (R87), and HO2 + CH3 ↔ OH + CH3O (R119). The negative
Figure 5.16 shows the radial distributions of calculated variables across the reaction
kernel (x = 0.33 mm). In general, the distributions of variables are similar to those at x
= -–0.1 mm, except for the increased levels of species production rates, net reaction
rates, and heat-release rates resulting in higher species concentrations and peak
species are primarily formed on the fuel-side and consumed on the air-side of the
stoichiometric mixture fraction; and the H, O, and OH radicals are produced on the
air-side and consumed on the fuel-side. The peak consumption rate of CH 4 is on the
fuel-side and that of O2 and the production rates of the H 2O and CO2 are on the air-
side. Figures 5.16(c) and (d) show that the chain-branching reaction (R38) is still the
most significant reaction and the methyl oxidation reaction (R10) is the major
contributor to the heat-release rate. It should be noted that the methyl (CH 3) radical is
produced more by the H atom through (R53) than by the OH radical through (R98) as
compared to those reactions at x = -–0.1 mm. Also, the formation (R168) and
destruction (R119) of HO2 reactions, that play an important role for the hot zone
formation in the gap region (x = -–0.1 mm), become insignificant at this location. The
dominant exothermic reactions, including R10, R284, R58, R290, R98, R99, R84,
R101, R125, and R168, contribute up to 81.5% of the total heat-release rate at the
reaction kernel (1.15 109 W/m3), thereby stabilizing a standoff microjet flame.
From the analysis of chemical kinetic structures, the sequence of flame stabilization
flame is quenched on the tube wall, heat is transferred through the tube wall to
accelerate fuel pyrolysis near the tube exit, produce CH3 and H intermediate radicals
in the standoff region, and initiate further reaction in the vicinity of the flame base.
reactions (R98) and (R53), and the final product formation (R84) and (R99) are
enhanced to further build up the radical pool of H, O, OH, and CH 3 in the gap region.
Then the methyl (CH3) radical oxidations (R10) and (R284) not only release
significant heat to enhance further reactions but also produce the formaldehyde
(CH2O) radical. The CH2O radical is attacked by the OH radical (R101) and H atom
(R58) to form the formyl (HCO) radical and followed by the HCO oxidation (R168)
to produce the HO2 radical. Finally the HO2 radical predominantly diffuses and moves
inward to the fuel stream near the tube exit and reacts with CH 3 (R119) to release heat
near the tube wall region. These key reactions are believed to form the hot zone (T =
800 ~ --1450 K) that connects the visible flame base and provides heat for sustaining
and enhancing further H2---–O2 chain reactions and CH3 formation (R53 and R98) and
oxidation (R10 and R284) at the flame base, which in turn result in the formation of
the reaction kernel responsible for flame stabilization. It should be noted that although
the reaction rate and heat-release rate of the HO 2 formation (R168) and consumption
(R119) reactions are not as significant as those of the other reactions;, they indeed
play an important role in the formation of the hot zone and high heat-release reaction
kernel [4135].
5.5 6 Summary
Flame characteristics, in terms of flame shape, flame length, flame structure, reaction
diffusion flames operated at various fuel exit velocityies ranging from just above
flame parameters, such as flame heights, flame shapes, and quenching velocities with
theoretical predictions indicate that Roper’s model [14] can satisfactorily predict the
velocity with measured results indicate that quenching occurs when the flame length
equals the standoff distance. It is also found that the quenching velocity for different
of the dimensionless parameters suggests that the microjet diffusion flames studied
here are not buoyant buoyancy-free and the molecular- diffusion is effective only for
NDB > 2 and Pe < 2. Numerical simulations of the flames stabilized at the tip of a 186
µm tube indicate that the computed flame shape and flame length are in excellent
show that the diffusion process dominates over the premixing process in the standoff
region, suggesting that the flame burns in a diffusion mode. Besides, the calculated
OH mass fraction isopleths indicate that the change of tube materials has a minor
effect on the standoff distance near extinction, but does influence the quenching gap
The computed flame heights as a function of fuel exit velocities indicate that
buoyancy may play a role for flames with larger tube diameters and higher exit
velocities. However, for the case of d = 186 µm flame with ue < 3.68 m/s, the
buoyancy effect becomes minor and the flame is diffusion controlled. Although
partial premixing may occur in the standoff region, its mixing intensity is not strong
enough to generate reaction and hence no double flame structure is observed. Further
examination of the computed flame structure reveals that the flame is stabilized by a
the hot zone through reaction of HO2 with CH3, yielding subsequent key radical
reactions, and finally forming the reaction kernel. The major reaction path near the
DoO :mean diffusion coefficient evaluated for the oxidizer at the oxidizer stream
temperature
f :focal length; mixture fraction
Fr :Froude number
fst :stoichiometric value of mixture fraction
gx :gravitational acceleration in the x-direction
h :enthalpy
I :ratio of the actual initial momentum to that for uniform flow
k :thermal conductivity
kij :thermal-diffusion ratio
lD :radial position of the flame edge
Le :Lewis number
Lf :flame length
M :molecular weight
NDB :diffusion to buoyancy parameter
NDM :diffusion to momentum parameter
p :pressure
Pe :Peclet number
Pi :power of incident laser
Qf :volumetric fuel flow rate
Re :Reynolds number
Ro0 :universal gas constant
S :molar stoichiometric oxidizer---fuel ratio
Sc :Schmidt number
T :temperature
u* :dimensionless axial velocity component
ud :molecular- diffusion velocity
ue :fuel exit velocity
w :species production rate
Wi :atomic weight of species i
x* :flame shape
xfl* :dimensionless flame height