You are on page 1of 8

Proceedings of the Combustion Institute, Volume 29, 2002/pp.

2587–2594

RADIATION-DRIVEN FLAME SPREAD OVER THERMALLY THICK FUELS IN


QUIESCENT MICROGRAVITY ENVIRONMENTS

YOUNGJIN SON and PAUL D. RONNEY


Department of Aerospace and Mechanical Engineering
University of Southern California
Los Angeles, CA 90089, USA

Microgravity experiments on flame spread over thermally thick fuels were conducted using foam fuels
to obtain low density and low thermal conductivity and thus large spread rate (Sf) compared to dense fuels
such as polymethylmethacrylate. This scheme enabled meaningful results to be obtained even in 2.2 s drop
tower experiments. It was found that, in contrast to conventional understanding, steady spread can occur
over thick fuels in quiescent microgravity environments, especially when a radiatively active diluent gas
such as CO2 is employed. This is proposed to be due to radiative transfer from the flame to the fuel surface
that can lead to steady spread even when conductive heat transfer from the flame to the fuel bed is
negligible. Radiative effects are more significant under microgravity conditions because the flame thickness
is larger and thus the volume of radiating combustion products is larger at microgravity. The effects of
oxygen concentration and pressure are shown and the transition from thermally thick to thermally thin
behavior with decreasing bed thickness is demonstrated. A simple semiquantitative model of radiation-
driven flame spread rates is consistent with experimental observations. Radiative flux measurements con-
firm the proposed effects of diluent type and gravity level. These results are particularly noteworthy con-
sidering that the International Space Station employs CO2 fire extinguishers; our results suggest that helium
may be a better inerting agent on both mass and mole bases at microgravity even though CO2 is much
better on a mole basis at earth gravity.

Introduction where dg is the length of the zone over which heat


is transferred from the gas to the fuel surface; for
It is well known [1–4] that convection influences opposed flow flame spread, dg is proportional to
flame spread over solid fuel beds in numerous ways. the convection-diffusion zone thickness ␣g/U
Flame spread is typically classified as opposed flow, where ␣g ⬅ kg/qgCp,g and U is the opposed flow
where the direction of flame propagation is opposite velocity. If self-induced convection is included,
convective flow past the flame front, or concurrent then dg ⳱ ␣g/(U Ⳮ Sf).
flow, where convection and spread are codirectional. For thermally thin fuel beds (where transverse
Downward spread at earth gravity (1g) is generally conduction within the bed is negligible), heat trans-
characterized by opposed flow since the upward fer occurs via gas-phase conduction; thus q ⬃
buoyant flow is opposite the direction of spread. At kg(Tf ⳮ Tv)/dg, where Tf is the flame temperature
quiescent microgravity (lg) conditions, flame spread [1] and thus
is necessarily opposed flow (unless a forced flow is
imposed) because the flame spreads toward the kg Tf ⳮ Tv
fresh atmosphere with self-induced convection ve- Sf ⳱ A (A ⳱ constant) (2)
qsCp,sss Tv ⳮ T⬁
locity equal to the spread rate (Sf). At 1g, self-in-
duced convection can justifiably be ignored since
buoyancy-induced flows are typically 10–30 cm/s, deRis [1] derived an approximate solution A ⳱ Z2
which is much higher than Sf; however, at lg self- and Delichatsios [5] found an exact solution A ⳱ p/4.
induced convection is dominant. For thermally thin fuels, steady spread is possible
Sf can be modeled by equating the heat flux per even at lg because this ideal Sf is independent of U.
unit area from the gas to the fuel surface (q) to the For thermally thin fuels, ss is the fuel bed half-
rate of increase of solid fuel bed enthalpy [2], leading thickness, whereas for thermally thick (effectively
to semi-infinite) fuels, where heat conduction through
the solid fuel is important, ss is the thermal penetra-
tion depth into the solid fuel (sp), estimated by
qdg equating q to the heat flux within the solid fuel,
Sf ⳱ (1)
qsCp,s (Tv ⳮ T⬁)ss ks[(Tv ⳮ T⬁)/sp]:

2587
2588 MICROGRAVITY COMBUSTION

ks (Tv ⳮ T⬁) lg. This was attributed to (1) increased radiative


sp ⳱ ⇒ emission from CO2 or SF6, which increases the net
q
heat flux to the fuel bed, and (2) reabsorption of this
q2dg radiation, which reduces radiative heat loss. (Diluent
Sf ⳱ (3) type also affects Lewis numbers but these effects
qsCP,s ks(Tv ⳮ T⬁)2
were shown to be less important.)
This result is identical to that derived by Tarifa and
Torralbo [6] and deRis [1], where q is a prescribed
externally imposed radiative source, so the present Approximate Analysis
approach is considered valid. If heat transfer to the
fuel bed occurs via conduction, q ⬇ kg(Tf ⳮ Tv)/dg When radiative heat transfer to the fuel bed is sig-
as for thin fuels, then the exact solution for Sf over nificant, Sf given by equations 1 and 3 is still valid
thick fuels [1] is obtained (neglecting self-induced but equation 4 must be modified. For flame-gener-
convection): ated radiation, qr is coupled to the spread process
itself. As a first estimate we consider optically thin
ksy dg(Tv ⳮ T⬁) radiation, where no reabsorption occurs and spectral
ss ⳱ ⇒
kg(Tf ⳮ Tv) properties are grouped into one parameter. We as-
2
sume the flame front to be an isothermal volume of
kg qgCp,g Tf ⳮ T
Sf ⳱ U 冢
ks qsCp,s Tv ⳮ T⬁ 冣 (4) optically thin radiating gas at temperature Tf with di-
mension dg in directions both parallel and perpendic-
ular to the fuel bed. We make this choice because for
The transition from thermally thin to thermally thick
optically thin radiation no length scale exists; thus the
behavior occurs when ss ⬇ sp [1]. A given material
thermal thickness of the flame is still determined by
may behave as thermally thin or thermally thick de-
the convective-diffusive zone thickness dg. Of course,
pending on U.
radiation can be transferred to the fuel bed from
Equation 4 shows that for thermally thick fuels,
distances greater than dg, but this radiation is at
Sf ⬃ U, suggesting that at lg, where U ⳱ 0, no
small angles to the fuel surface and thus is not an
steady spread is possible. If self-induced convection
effective heat transfer mechanism. The heat flux due
(U ⳱ Sf) is considered, then Sf is indeterminate un-
to radiation can be estimated as Kdg, where K ⳱
less forced flow is applied to generate some U in-
4raP(Tf4 ⳮ Tv4) is the radiation power per unit vol-
dependent of Sf. An unsteady analysis for quiescent
ume. The combined effects of gas-phase radiation and
thermally thick flame spread at lg [7] predicts that
thermal conduction is q ⳱ Kdg Ⳮ kg(Tf ⳮ Tv)/dg.
ss ⬃ (␣st)1/2, which results in Sf ⬃ tⳮ1/2, where t is
Combining this with dg ⳱ ␣g/Sf and equation 4 leads
the time lapse from ignition. The prevailing opinion
to, assuming unit fuel bed emissivity and U ⳱ 0,
[7] is that Sf should decrease continually until ex-
tinction occurs via radiative losses. Computations 1/2
K␣ g2
and space experiments [7,8] supported these asser-
tions.
Sf ⳱ 冤冪␣ q C
g s P,s ks (Tv ⳮ T⬁) ⳮ kg (Tf ⳮ Tv)

In this work, we show that this inability to obtain (5)
steady spread over thermally thick fuels in quiescent
Equation 5 predicts that without gas-phase radia-
lg environments no longer applies when heat trans-
tion, no steady spread is possible (Sf ⳱ 0) and with
fer to the fuel bed via flame-generated radiation (in
gas-phase radiation, Sf ⬃ K1/2. Thus, increasing gas-
addition to or instead of conduction) is significant.
phase radiation should increase Sf. Of course this
Steady spread could also result from an externally
also increases radiative loss, but the ratio of heat loss
imposed radiative source (equation 3) and has been
to heat generation will remain roughly constant.
demonstrated previously [9]. Altenkirch and collab-
Equation 5 also shows that pressure effects are im-
orators [7,8] utilized systems where flame radiation
portant since K ⬃ P and ␣g ⬃ Pⳮ1.
acted primarily as a loss mechanism. We derive an
Radiative transfer to the fuel bed will also increase
approximate expression for Sf and conduct lg ex-
Tf [1], although using representative values of the
periments that demonstrate the proposed mecha-
thermodynamic and transport properties the pre-
nism of flame spread with flame-generated radiation.
dicted effect is too weak to affect the above conclu-
As evidence of the importance of this flame radi-
sions. It does, however, make the impact of radiation
ation, thin-fuel flame spread experiments [10] in ra-
slightly stronger than predicted here.
diatively inert N2, He, or Ar diluents showed the
conventional behavior where Sf is lower and the min-
imum flammable O2 concentration is higher at lg. Experimental Apparatus
This occurs because when U is lower, dg is larger and
the volume of radiating gases is higher. In contrast, To investigate the possibility of steady flame
for radiatively active CO2 and SF6 diluents, the op- spread over thermally thick fuels in quiescent lg en-
posite behavior was observed where Sf is higher at vironments, experiments were conducted in the
RADIATION-DRIVEN FLAME SPREAD 2589

NASA-Glenn 2.2 s drop tower and downward prop- conductivity (ks) and density (qs) than PMMA. Poly-
agating (opposed-flow) comparison tests were per- phenolic foams were chosen primarily because they
formed at 1g. Experiments were performed in our have lower sooting tendency and negligible melting
existing flame spread apparatus [10]. The 20 L or dripping tendency compared to other foams such
chamber was filled with desired atmospheres by par- as polystyrene or polyurethane. Of course, all foams
tial pressures. The fuel samples (described below) contain trapped gas; however, the density of the foams
were 10 cm wide ⳯ 11.5 cm long and were held employed is at least 20 times larger than the test at-
between aluminum quenching plates to inhibit edge mospheres. Thus, the trapped gas contributes negli-
burning. This sample width yielded nearly width-in- gibly to the overall stoichiometry. The permeability of
dependent Sf ; thus three-dimensional effects were the foam (typically 10ⳮ7 m2) is small enough that flow
probably not significant. Unless otherwise noted, through the porous medium is negligible. Equation 4
samples were 10 mm thick, which we shall show is also shows that high pressure (thus high qg) favors
well into the thermally thick limit. Both single-sided high Sf and was employed here.
and two-sided burning was examined; unless other-
wise noted, all results shown were for single-sided
burning. After allowing time for settling of convec- Experimental Results
tion, the fuel samples were ignited by an electrically
heated Kanthal wire imbedded in a nitrocellulose It was found that by using foam fuels, position ver-
membrane (energy release ⬎1000 J) glued onto the sus time plots (not shown due to space limitations)
fuel surface. Ignition was controlled and radiometer usually exhibited statistically straight lines (not
data are collected by a microcomputer on the drop slowly bending) typically 1 s after ignition, indicating
package. The flames were imaged using CCD cam- steady spread even in these short-duration lg ex-
eras whose signals were connected via fiber-optic ca- periments. Some tests at low pressures and/or low
bles to ground-based video recorders. One camera oxygen concentrations neither extinguished nor
was positioned with its viewing axis perpendicular reached steady state within 2.2 s of lg time and were
the fuel sample plane to images the flame front and discarded. While we cannot positively rule out tran-
thus spread rate. Another camera imaged laser sitions to different steady spread rates or to extinc-
shearing interferograms of the flames from a side tion at longer times for longer samples, there was no
view to compare qualitatively flame shape and thick- evidence of this. Also, during the lg tests our flames
ness at 1g versus lg. spread much farther than dg estimated both theo-
Narrow-angle thermopile-type radiometers retically and from the interferometer images, mean-
mounted 10 cm from the fuel bed were used to de- ing that there is little likelihood of such transitions.
termine radiative emissions from the flames. The Fig. 1 (top) shows that, as was also seen in the thin-
measured radiative fluxes represent only the flux fuel tests [10] and in ETFE wire burning tests [11],
reaching the detector through the narrow-angle ra- for CO2-diluted atmospheres, Sf could be higher at
diometer and thus are much smaller than the total lg than 1g, especially at low O2 concentrations, but
emissive power from the flame. Three radiometers for N2-diluted atmospheres Sf was always higher at
were used: (1) a front-side (burning side) radiometer 1g than lg. At lg, Sf can actually be higher in CO2
viewing a hole in the fuel bed to measure only the than N2 at the same O2, concentration even though
outward gas-phase radiative loss, (2) another front- CO2 has a larger CP and thus yields lower Tf than
side radiometer viewing the fuel surface to measure N2 for the same O2 concentration.
both the outward gas-phase and surface radiative Figure 1 (top) also shows that for CO2 diluent, the
fluxes and thus total radiative loss, and (3) a back- minimum O2 concentration supporting combustion
side radiometer viewing through the hole to mea- is substantially lower at lg than 1g, whereas for N2,
sures the inward gas-phase radiative heat flux and the minimum O2 concentration is higher at lg. All
thus the fuel bed heating due to gas-phase radiation. of these results show that flames in CO2-diluted at-
The difference between (2) and (1) is the surface mospheres burn more robustly at lg than 1g,
contribution to the total radiative loss. whereas the opposite trend is found for N2. We pro-
The standard fuel for fundamental thick-fuel pose that this is due to three factors. First, K is larger
flame spread experiments is polymethylmethacrylate for O2/CO2 atmospheres (since both the combustion
(PMMA) which has the thick fuel spread rate param- products and the ambient atmosphere contain radi-
eter ksqsCP,s(Tv ⳮ T⬁)2 ⬇ 3.3 ⳯ 1010 J2/m4s. ant species, whereas for O2/N2, only the combustion
This leads to Sf ⬇ 0.006 cm/s in air at 1 atm, which products radiate) which increases the heat flux to the
is too low to observe steady-state spread in short-du- fuel bed and thus Sf (equation 5). Second, without
ration drop tower experiments. After evaluating nu- buoyant convection (U ⳱ 0), the flame thickness
merous candidate materials, we chose open-cell dg ⳱ ␣g/(U Ⳮ Sf) is much thicker at lg than at 1g;
polyphenolic foams that have values of ksqsCP,s thus, lg flames have more volume and can transfer
(Tv ⳮ T⬁)2 2 to 3 orders of magnitude smaller than more radiation to fuel bed. Interferometer images
PMMA because the foams have much lower thermal (not shown) confirm that flames are much thicker at
2590 MICROGRAVITY COMBUSTION

lg. This effect is more important for lower O2 con-


centrations (thus lower Sf) which explains why the dif-
ference between 1g and lg spread rates in O2/CO2
atmospheres is larger at lower O2 concentrations.
Third, O2/CO2 atmospheres can reabsorb and re-
radiate emitted radiation, whereas O2/N2 atmo-
spheres cannot; thus substantial radiative heat losses
that would otherwise occur at lg with thick flames
in strongly radiating O2/CO2 atmospheres are at
least partially suppressed.
The difference between CO2- and N2-diluted at-
mospheres is less at higher O2 concentrations, prob-
ably because at higher O2 reactant concentrations
the CO2 and H2O product concentrations are higher,
thus increasing the effective K for O2/N2 atmo-
spheres having no radiatively participating specie. In
contrast, for O2/CO2 atmospheres increasing O2 de-
creases ambient CO2 concentrations, thus decreas-
ing K.
Figure 2 (top) shows pressure effects on Sf. Except
at low pressures, for CO2 diluent, Sf is higher at lg
than at 1g, whereas for N2 diluent, only at very high
Fig. 1. Effects of oxygen concentration on spread rates
pressures is Sf higher at lg. The reversal of CO2
over thick solid fuel beds at lg and at 1g, polyphenolic
behavior at low pressures may occur because flame
foam, density 0.0267 g/cm3, 4 atm total pressure. (Top)
thickness is larger at low pressure since ␣g is larger
Comparison of O2/N2 and O2/CO2 mixtures; (bottom) and because absorption coefficient aP is smaller at
comparison of O2/He and O2/CO2 mixtures. lower pressure; thus at sufficiently low pressure, ra-
diative loss may dominate reabsorption and reradi-
ation effects. Sf increases with increasing pressure
until a plateau of Sf is reached for both lg and 1g
cases. This transition may be due to transition to
nearly opaque radiative conditions (aPⳮ1 ⬍ dg) since
aP is larger at higher pressure. Fig. 2 (bottom) shows
predicted pressure effects on Sf (equation 5) for
O2/CO2 atmospheres at lg. Two different assumed
values of Tf are shown. All gas properties are eval-
uated at the average temperature (Tf Ⳮ T⬁)/2. The
model provides reasonable Sf estimates except near
the low-pressure extinction limit, where heat losses
may dominate, leading to slower than predicted
flame spread, and at high pressure where there may
be a transition to nearly opaque conditions. The
opacity at high pressure causes less radiative transfer
thus less fuel bed preheating than predicted by our
optically thin model, which leads to lower Sf than
the optically thin predictions. (Neither of these ef-
fects can be predicted by the simple optically thin,
loss-free model leading to equation 5).
Figure 3 shows the effects of fuel bed thickness
on spread rate for a fixed atmosphere. Sf is indepen-
dent of thickness for sufficiently thick fuel beds,
Fig. 2. Effects of pressure on spread rate over thick solid demonstrating thermally thick conditions. The thick-
fuel beds at lg and 1g, polyphenolic foam, density 0.0267 ness at the start of the transition is about 2 mm for
g/cm3. (Top) 40% O2/60% CO2 and 40% O2/60% N2, 1g the cases shown. For thick samples, Sf is nearly the
and lg. (Bottom) 40% O2/60% CO2 at lg only, along with same for one-sided and two-sided burning, again in-
spread rate estimates based on equation 5 assumed flame dicating thermally thick flame spread. For the thin-
temperatures of 1500 and 1800 K. In these estimates, all nest samples, Sf for one-sided spread is about half
gas properties are calculated using temperature averaging that of two-sided spread, which is consistent with the
between ambient temperature and flame temperature. simple thermal model for thin fuels (equations 1 and
RADIATION-DRIVEN FLAME SPREAD 2591

at lg (Fig. 4a). This is probably because only in this


case is there substantial emission, absorption, and
re-emission, which is the only means to obtain sub-
stantial radiative flux to the back-side radiometer.
O2/N2 atmospheres do not show this behavior at all,
and even for O2/CO2 atmospheres, this is seen only
at lg where dg is larger and thus the total radiative
flux ⬃Kdg is greater. Note also that the gas-phase
radiative loss (‘‘front, gas only’’ curves) at lg is ac-
tually lower for O2/CO2 than O2/N2 atmospheres
due to reabsorption by the ambient atmosphere for
O2/CO2. At 1g (Figs. 4b and 4d), the surface radia-
tion (the difference between the two front-side ra-
diometer readings) is much larger than gas-phase ra-
Fig. 3. Effects of fuel bed thickness on spread rate over diation due to the decreased flame thickness thus
solid fuel beds at lg and 1g, polyphenolic foam, density decreased volume of radiating gas at 1g. These re-
0.0267 g/cm3, 40% O2/60% CO2, 4 atm total pressure. sults confirm our hypotheses concerning radiative
transfer effects on lg flame spread, in particular that
(1) radiative preheating of the fuel bed by the gas is
2). Note however the odd behavior that Sf actually significant in radiatively active atmospheres at lg, (2)
decreases with decreasing thickness, which is con- reabsorption effects can prevent massive heat losses
trary to thin-fuel theory without radiation (equation (thus extinction) in radiatively active atmospheres at
2) and will never occur without radiation because at lg, and (3) these effects are less important at 1g due
the transition from thermally thin to thermally thick to substantial U caused by buoyancy which leads to
behavior (sp ⳱ ss), the values of Sf are equal for thin smaller flame thicknesses and thus less volume of
(equation 2) and thick (equation 4) fuels. With ra- radiating gas.
diative effects, this is no longer true; Sf for thick
fuels is given by equation 5, and for thin fuels Sf is
determined from equation 1 using q ⳱ Kdg Ⳮ Discussion and Conclusions
kg(Tf ⳮ Tv)/dg and dg ⳱ ␣g/(U Ⳮ Sf), leading to a
cubic equation for Sf: Microgravity experiments on flame spread over
3 2 thermally thick fuels were conducted using foam fu-
冢SS 冣
f

f,o
Ⳮ 冢S2U ⳮ 1冣冢SS 冣
f,o f,o
f
Ⳮ 冢
U U
Sf,o Sf,o
ⳮ2冣冢 冣
Sf
Sf,o
els to obtain low density and thermal conductivity
and thus large spread rate (Sf) compared to dense
Sf,rad 3 fuels such as PMMA. This scheme enabled mean-

ⳮ U2 Ⳮ 冢S 冣冣⳱0
f,o
(6) ingful results to be obtained even in 2.2 s drop tower
experiments. It was found that, in contrast to con-
where Sf,o is the radiation-free spread rate (equation ventional understanding, steady spread could occur
2) and Sf,rad ⳱ [K␣g2/qsCP,sss(Tv ⳮ T⬁)]1/3 is the over thick fuels in quiescent lg environments, es-
spread rate without conduction to the fuel bed. (The pecially when a radiatively active diluent gas such as
effect of U for thick fuels can be computed in the CO2 is employed. In some cases with CO2 diluent,
same way, but the resulting fourth-order polynomial the spread rate was actually higher at lg than at 1g
is too long to print here.) It can be shown that at the despite the absence of convection at lg, which with-
transition point sp ⳱ ss, the thin fuel Sf (equation 6) out radiative transfer is expected to preclude the
can be lower than the thick fuel Sf (equation 5), lead- possibility of steady spread. This was shown to be
ing to non-monotonic effects of ss on Sf. Moreover, due to radiative transfer from the flame to the fuel
this behavior can occur with or without an imposed surface. This assertion is consistent with measure-
or buoyant flow U; thus, the behavior can be seen at ments of the radiatively fluxes to and from the fuel
both 1g and lg. (The foams could not be sliced thin bed. This conclusion was also supported by interfer-
enough to observe conventional thin-fuel behavior ometer images showing that the flames where much
at lg (Sf increasing as ss decreases) but this behavior thicker at lg than 1g, indicating that the lg flames
is seen for 1g flame spread in Fig. 3.) have more volume and thus can radiate more heat
Figures 4a–4d show the radiative characteristics of to the fuel bed. Additionally, the transition from
O2/CO2 and O2/N2 flames at 1g and lg. The only thermally thick to thermally thin behavior with de-
case where the back-side radiometer (which mea- creasing bed thickness was demonstrated, at a typical
sures the gas-phase radiant heat flux to the fuel bed; fuel bed thickness of 2 mm.
see Experimental Apparatus) shows comparable in- These results are relevant to fire safety in manned
tensity and timing with the two front-side radiome- spacecraft, particularly the International Space Sta-
ters is for the radiatively active O2/CO2 atmosphere tion that uses CO2 fire extinguishers. CO2 may be a
2592 MICROGRAVITY COMBUSTION

a b

c d
Fig. 4. Radiative flux characteristics of flames spreading over thick polyphenolic foam fuel. See Experimental Apparatus
section for discussion of the interpretation of radiometer data. (a) 40% O2/60% CO2, 4 atm, lg. (b) 40% O2/60% CO2,
4 atm, 1g. (c) 40% O2/60% N2, 4 atm, lg. (d) 40% O2/60% N2, 4 atm, 1g.

less effective inerting agent at lg than at 1g because whereas at lg the minimum O2 mole fraction was
of the differences in spread mechanisms at 1g and much higher for He (35%) than CO2 (27%). There
lg. In particular, the difference between conduc- are at least three reasons for the observed behavior.
tion-dominated heat transport at 1g versus radiation- First, the Lewis number of O2 in He is much higher
dominated heat transport at lg indicates that radia- than O2 in CO2 (⬇1.20 vs. ⬇0.84), which leads to
tively inert diluents such as helium could be lower spread rates for He [12]. Second, the higher
preferable in lg applications. With this motivation, kg and ␣g of He leads to thicker flames and thus
tests were conducted using He diluent. Fig. 1 (bot- greater radiative loss for the same Sf. Third, unlike
tom) shows that He- and CO2-diluted atmospheres CO2, He is radiatively non-participating and thus no
exhibit nearly the same Sf for a fixed O2 mole fraction reabsorption or re-emission occurs. Consequently,
we conclude that He may be a superior inerting
even though CO2 has a mole-based CP (at 300 K)
agent at lg on several bases. First, at lg. He is more
1.8 times higher than He and (at 2000 K) 3.9 times effective than CO2 on a mole basis (thus pressure
higher. Also, He has a thermal conductivity (kg) 9.4 times storage volume basis), meaning that the size
times higher than CO2. Both of these factors should and weight of storage bottles would be smaller for
lead to higher Tf and Sf in He- than CO2-diluted the same fire-fighting capability. Second, He is much
atmospheres at the same O2 concentration (equa- more effective on a mass basis (by about 11⳯) at
tions 3 and 5). Furthermore, at 1g, the minimum O2 lg. Third, He has no physiological activity, unlike
mole fraction supporting combustion was nearly the CO2, which affects human respiration. Fourth, com-
same (30%) in He- and CO2-diluted atmospheres, pared to N2 or CO2, He is not very soluble in water
RADIATION-DRIVEN FLAME SPREAD 2593

and thus has less tendency to cause bloodstream Gokoglu, Dr. Linton Honda, and the NASA-Glenn 2.2 s
bubble formation following spacecraft cabin depres- drop tower staff for helpful discussions and technical sup-
surization. port.

Nomenclature
aP Planck mean absorption coefficient
CP constant-pressure heat capacity REFERENCES
qr radiant heat flux per unit area
1. deRis, J. N., Proc. Combust. Inst. 12:241 (1968).
qk radiant heat flux per unit wavelength
2. Williams, F. A., Proc. Combust. Inst. 16:1281 (1976).
Sf flame spread rate over solid fuel bed
3. Fernandez-Pello, A. C., Combust. Sci. Technol. 39:119
T temperature
(1984).
U convection velocity
4. Wichman, I. S., Prog. Energy Combust. Sci. 18:553
␣ thermal diffusivity
d flame thickness (1992).
k thermal conductivity 5. Delichatsios, M. A., Combust. Sci. Technol. 44:257
K radiative heat transfer per unit volume (1986).
q density 6. Tarifa, C. S., and Torralbo, A. M., Proc. Combust. Inst.
ss fuel bed half-thickness (thin fuel) 11:533–544 (1966).
sp thermal penetration depth (thick fuel) 7. Bhattacharjee, S., West, J., and Altenkirch, R. A., Proc.
r Stefan-Boltzman constant Combust. Inst. 26:1477 (1996).
8. Altenkirch, R. A., Tang, L., Sacksteder, K., Bhatta-
Subscripts charjee, S., and Delichatsios, M. A., Proc. Combust.
f flame front condition Inst. 27:2515 (1998).
g gas-phase condition 9. Brebob, E. G., and Kulkarni, A. K., Fire Materials
s solid fuel or solid surface condition 17:249 (1993).
v solid fuel vaporization condition 10. Honda, L. K., and Ronney, P. D., Combust. Sci. Tech-
⬁ ambient conditions nol. 133:267 (1998).
11. Fujita, O., Kikuchi, M., Ito, K., and Nishizawa, K.,
Acknowledgments
Proc. Combust. Inst. 28:2905 (2000).
This work was supported by NASA-Glenn Research 12. Zhang, Y., Ronney, P. D., Roegner, E., and Greenberg,
Center under grant NCC3-671. We thank Dr. Suleyman J. B., Combust. Flame 90:71 (1992).

COMMENTS
Takashi Kashiwagi, NIST, USA. I am extremely sympa- spread rate without adjustable parameters. Second, the
thetic that you were forced to conduct your experiments spread rate was sometimes higher with CO2 than N2 dilu-
at a high oxygen concentration (40%) atmosphere because ent at the same O2 concentration, even though (as men-
of a limited experimental time. At such a high oxygen con- tioned in the paper) CO2 diluent should produce lower
centration, flame is extremely bright, and gas-phase radi- flame temperatures. Third, based on the visible images, it
ation becomes very significant. However, in air, the flame appeared that the soot formation was lower in CO2 than
tends to be blue in microgravity, and gas-phase radiation N2-diluted atmospheres, even at the same O2 concentra-
may not be as dominant as in 40% O2. What is your esti- tion. As you mentioned, at lower O2 concentrations, the
mate of the contribution of the gas-phase radiation transfer flame is more blue and less yellow, indicating less impor-
in CO2 atmosphere to flame spread in air that is the main tance of soot formation, and thus we would expect an even
atmosphere in the space station? higher contribution of gas-phase radiation relative to soot
formation. For thermally thick fuels in a quiescent micro-
Author’s Reply. The brightness you mention is due to gravity environment, without the gas-phase radiative feed-
soot formation, since CO2, CO, and H2O do not radiate back, there is no means known to us to support steady
significantly in the visible spectrum. Thus, one should not flame spread, and if radiative losses are not sufficient to
judge the magnitude of gas-phase radiation based on visible extinguish the flame, according to our model, steady spread
emission. Our model does not include the effects of soot is always possible—although the flame dimension and
radiation. While we have not quantified the relative im- timescale will be very large if the radiation is weak as may
portance of soot versus gas-phase radiation (e.g., via spec- be the case in air at 1 atm.
trally resolved emission measurements), there are several

facts that suggest gas-phase radiation dominates our re-
sults. First, our simple analytical model that does not in- Osamu Fujita, Hokkaido University, Japan. Spread rate
clude soot radiation gives semiquantitative predictions of with N2 dilution in microgravity is almost the same as in
2594 MICROGRAVITY COMBUSTION

1g. It means the condition is more like a thermal regime thick regime because the spread rate is independent of fuel
rather than a radiation control regime. How do you explain bed thickness (Fig. 3). Therefore, a spread rate indepen-
this? dent of the opposed flow velocity (e.g., the same spread
rate at 1g and microgravity) indicates that something other
Author’s Reply. Your suggestion that equal spread rates than conduction (i.e., radiation) is controlling the heat
means the thermal regime (which we interpret to mean transport to the fuel bed, and thus the thermal regime is
heat transfer to the fuel bed dominated by conduction) is not operative. It is also important to note that only for some
operative would be appropriate for thermally thin fuels, conditions, for example, higher pressures, are the 1g and
where the ideal spread rate is independent of the opposed microgravity spread rates similar. At lower pressures, the
flow velocity. In contrast, for thermally thick fuels, in the 1g spread rates are much higher than the microgravity
thermal regime (considering conduction but not radiation spread rates, which is more indicative of the thermal re-
to the fuel bed) the spread rate is proportional to the op- gime where spread rate is proportional to opposed flow
posed flow velocity. We know that we are in the thermally velocity for thermally thick fuels.

You might also like