ArticlePDF Available

Smoldering propagation and blow-off on consolidated fuel under external airflow

Authors:

Abstract and Figures

Propagation of smoldering combustion and its blow-off phenomena are of practical importance in evaluating the fire dynamics of solid fuels, but the scientific understanding is still limited. In this work, we quantify the smoldering propagation rates on consolidated biomass and the blow-off limits under concurrent and opposed external airflows up to 50 m/s. The incense cylinders with different diameters (1.5-5.0 mm) and densities (720-1,100 kg/m 3) are tested. As the airflow velocity increases, the smoldering propagation rate first increases to its maximum value (Oxygen-limited Regime) and subsequently remains stable (Thermal Regime), regardless of the airflow direction. Afterward, it slightly decreases (Chemical Regime) until blow-off, and the blow-off of opposed smoldering is easier, similar to the pattern of flame spread. The blow-off airflow velocity (13-46 m/s) of smoldering combustion is around ten times larger than that of flaming combustion, and it decreases as the fuel diameter or density increases. This work advances the fundamental understanding of the smoldering propagation, blow-off, and its persistence; thus, helping guide the fire suppression strategies of smoldering.
Content may be subject to copyright.
S. Lin, T. Chow, X. Huang (2021) Smoldering Propagation and Blow-off on Consolidated Fuel under External Airflow,
Combustion and Flame, 234, 111685. https://doi.org/10.1016/j.combustflame.2021.111685
1
Smoldering Propagation and Blow-off on Consolidated Fuel
under External Airflow
Shaorun Lin1,2, Tsz Him Chow1, and Xinyan Huang1,*
1Research Centre for Fire Safety Engineering, The Hong Kong Polytechnic University, Kowloon,
Hong Kong SAR
2The Hong Kong Polytechnic University Shenzhen Research Institute, Shenzhen, Guangdong, China
*Corresponding to xy.huang@polyu.edu.hk (X. Huang)
Abstract
The propagation of smoldering combustion and the blow-off limit are of practical importance in
evaluating the fire dynamics of solid fuels, but the scientific understanding is still limited. In this work,
we quantify the smoldering propagation rate on consolidated biomass and the blow-off limit under
concurrent and opposed external airflows up to 50 m/s. The incense cylinders with different diameters
(1.5-5 mm) and densities (720-1,100 kg/m3) are tested. As the airflow velocity increases, the smoldering
propagation rate first increases to its maximum value (Oxygen-limited Regime) and subsequently
remains stable (Thermal Regime), regardless of the airflow direction. Afterward, it slightly decreases
(Chemical Regime) until blow-off, and the blow-off of opposed smoldering is easier, similar to the
pattern of flame spread. The blow-off airflow velocity (13-46 m/s) of smoldering combustion is around
ten times larger than that of flaming combustion, and it decreases as the fuel diameter or density
increases. This work advances the fundamental understanding of the smoldering propagation, blow-off,
and its persistence; thus, helping guide the fire suppression strategies of smoldering.
Keywords: smoldering fire; extinction limit; oxygen supply; biomass; wind effect.
1. Introduction
Smoldering is the slow, low-temperature, and flameless burning of porous fuels and one of the
most persistent types of combustion phenomena [13]. Smoldering combustion is a heterogeneous
process sustained when oxygen directly attacks the hot fuel surface, different from the flame regarding
the combustion chemistry and transport processes [2,3]. Smoldering can be ignited easily by a weak
heat source [24] or even self-ignited, which usually occur in silos and large fuel piles [5], creating a
shortcut to more intensive flaming fires (through smoldering-to-flaming transition). Moreover, it is also
challenging to detect and suppress the hidden smoldering fire. For example, the colossal piles of World
Trade Center debris continued to smolder for more than half a year, despite substantial firefighting
operations [6]. Natural smoldering, such as the underground fires in peatlands or coal mines, is one of
the most extensive and longest-lasting fire phenomena on Earth [7,8]. Therefore, it is vital to deepen
our understanding of smoldering fire dynamics.
S. Lin, T. Chow, X. Huang (2021) Smoldering Propagation and Blow-off on Consolidated Fuel under External Airflow,
Combustion and Flame, 234, 111685. https://doi.org/10.1016/j.combustflame.2021.111685
2
The fire spread (propagation) process is of practical significance in evaluating the impact of fire
events [912]. The fire spread is a continuous ignition and burning process [13], depending on both
environment (e.g., wind [10,11,1416], oxygen [1719], pressure [20,21], temperature [22,23], and
gravity [24]) and fuel factors (e.g., type/array [25], moisture [16], density [26], orientation [27], and
size [28,29]). Based on the relative direction to the airflow (or wind), fire spread can be classified into
the concurrent and opposed modes [911]. In the literature, most studies have focused on the
characteristics of flame spread on solid fuels [911,30], rather than the smoldering spread.
Smoldering combustion is controlled by the competition between the oxygen supply and the heat
transfer to and from the reaction zone [3,31,32]. Therefore, the airflow or wind is crucial to smoldering
propagation, because it could increase both the oxygen supply and the heat loss [15,29,33]. By applying
an external airflow (or environmental wind), smoldering propagation may become faster because of the
increased oxygen supply (O2-limit regime) [1,11,17]. Afterward, the excessive airflow may also help
Nomenclature
Symbols
Greeks
a
strain rate (s-1)
thermal diffusivity (m2/s)
c
specific heat capacity (J/kg-K)
δ
thickness (m)
d
fuel diameter (mm)
v
stochiometric coefficient (-)
C
fitting coefficient (-) / constant (-)
kinematic viscosity (m2/s)
D
wind tunnel diameter (m)
ρ
density (kg/m3)
Da
Damkohler number (-)
λ
thermal conductivity (W/m-K)
h
convection coefficient (W/m2-K)
ϕ
porosity (-)
hm
mass transfer coefficient (kg/m2-s)
reaction rate (1/s)
h
thermal enthalpy difference (J/kg)
Hc
heat of combustion (MJ/kg)
I
thermal inertia (J/m2-K-s1/2)
Subscripts
k
permeability (m2)
a
airflow/ambient
L
length (m)
ch
chemical
m"
mass flux (kg/m2-s)
con
concurrent
P
pressure difference (Pa)
cond
conduction
q"
heat flux (kW/m2)
conv
convection
regression rate (m/s)
ex
extinction
Re
Reynolds number (-)
f
fire
t
time (s)
F
fuel
T
temperature (K)
o
initial
ua
internal airflow velocity (m/s)
ox
oxygen
Ua
external airflow velocity (m/s)
p
preheating
V
fire propagation rate (cm/min)
r
residence

distance (m)
sm
smoldering
Y
mass fraction (%)
T
thermal
S. Lin, T. Chow, X. Huang (2021) Smoldering Propagation and Blow-off on Consolidated Fuel under External Airflow,
Combustion and Flame, 234, 111685. https://doi.org/10.1016/j.combustflame.2021.111685
3
trigger gas-phase homogenous oxidation under some specific conditions and result in smoldering-to-
flaming (StF) transition [19,34]. However, for flaming fires, flame spread increases with wind speed
due to increased convective heating on the unburnt fuel, rather than increased oxygen supply [33]. On
the other hand, the porosity and permeability of fuel also affect the oxidation-controlled smoldering
processes. For high-permeability fuels, such as cotton [15,35], pine needle [36], and PU foam [37],
oxygen can diffuse into the porous fuel to maintain an internal smoldering propagation. For low-
permeability consolidated fuels like wood [38], fiberboard [1], and coal chunk [8], smoldering can only
propagate from outside to inside like a regression process, because oxygen could only diffuse through
the porous char that is produced from the first-stage pyrolysis process [11]. Further increasing the
airflow velocity, the cooling effect becomes dominant, so eventually, smoldering extinction or blow-
off will occur, just like the blow-off of flame [39].
In the literature, the blow-off of flame on solid fuels has been extensively studied over the last 50
years [4,11]. For example, Loh and Fernandez-Pello [40] showed that the concurrent rate flame spread
over the thin paper first increased with the airflow velocity (< 1 m/s) but became almost constant until
blow-off at about 3 m/s. A similar trend and blow-off wind speed were also observed for the concurrent
flame spread on thin electrical wires [41]. In general, the blow-off of opposed flame spread is easier,
usually at an airflow velocity lower than 1 m/s [42,43]. Comparatively, the research on the blow-off of
smoldering is limited; and generally, it is more difficult to blow off persistent smoldering fire. Palmer
[1] found that the blow-off limit of opposed smoldering propagation over fiberboard was about 7 m/s,
but the concurrent smoldering propagation could still be sustained at 10 m/s [1,11]. Like the flame,
most smoldering extinction processes result from a local energy imbalance, where the cooling rate is
larger than the heat-release rate from exothermic oxidations [4,39,44]. Thus, decreasing oxygen
concentration and pressure promotes the blow-off of smoldering under a smaller airflow [19,20]. So far,
no study has addressed the smoldering propagation at large wind speeds over 10 m/s and the blow-off
limits of persistent smoldering fire; thus, there is a big knowledge gap.
This work investigated both concurrent and opposed smoldering propagations over cylindrical
consolidated biomasses (incenses) with different fuel diameters (1.5-5 mm) and densities (720-1,100
kg/m3). The external airflow velocity of up to 50 m/s in a small wind tunnel was applied to explore the
blow-off limits. The theoretical analysis was proposed to explain the influence of environmental and
fuel properties on smoldering propagation and critical conditions of blow-off.
2. Experimental Methods
2.1. Materials
The cylindrical consolidated rod (i.e., incense), a representative biomass fuel that is prone to
smoldering combustion, was tested in this work (Fig. 1a). The incense is an aromatic biotic material
that is widely used in cultural and religious events in Asia. It mainly consists of mixed wood dust from
the aromatic plants (e.g., from sage and cedar) and has homogenous porosity and composition [45]. The
S. Lin, T. Chow, X. Huang (2021) Smoldering Propagation and Blow-off on Consolidated Fuel under External Airflow,
Combustion and Flame, 234, 111685. https://doi.org/10.1016/j.combustflame.2021.111685
4
thermal analysis (TGA-DSC) of this incense was conducted, and the data is shown in Fig. A1 of the
Appendix. The details of the front and cross-section of the incense are also shown in Fig. 1a. Unlike the
cotton bales and plastic foams, the dust particles inside the incense are densely packed, so oxygen is
difficult to flow or diffuse into its internal structure.
Before the test, the incenses were first oven-dried at 75 oC for at least 48 h. Afterward, all samples
were placed into an electronic dry cabinet to avoid the re-absorbing of moisture from the air. To explore
the effect of fuel diameter () and density () on the smoldering propagation, two groups of experiments
were designed:
(Ⅰ) three sample diameters of 1.5, 2.5, and 5.0 mm with a constant fuel density of 720 kg/m3, and
(Ⅱ) three sample densities of 720, 920, and 1,100 kg/m3 with a constant diameter of 1.5 mm.
To help estimate the rate of smoldering propagation, the long incense rod was cut into 10-15 cm samples
and marked like a ruler with an interval of 1 cm (see Fig. 1a).
2.2. Environmental control
The experiments of smoldering propagation and blow-off under external airflow were conducted
inside a small wind tunnel. The customized tubular wind tunnel was made of quartz glass and had an
inner diameter () of 2 cm and a length of 20 cm, as illustrated in Fig. 1(b). The airflow (20.9% oxygen)
from the compressed tank was fed through the bottom of the quartz glass tube, and then homogenized
through a layer of small steel beads. A similar setup was used previously to study the flame spread [24]
and smoldering propagation [19] under opposed flow with different oxygen mass fractions. Before the
test, the airflow velocity ( up to 50 m/s) was controlled and measured by a precision anemometer.
Fig. 1. (a) Photos of cylindrical incenses with different diameters with enlarged details of surface and
cross-section, and (b) schematic of experimental setups for concurrent and opposed smoldering
propagation under external airflow.
S. Lin, T. Chow, X. Huang (2021) Smoldering Propagation and Blow-off on Consolidated Fuel under External Airflow,
Combustion and Flame, 234, 111685. https://doi.org/10.1016/j.combustflame.2021.111685
5
For an internal flow in a circular tube of diameter , the Reynolds number () can be calculated
as 
, where  m2/s is the kinematic viscosity of the air at ambient
temperature [46]. In other words, when the airflow velocity is larger than 2 m/s, the mainstream flow
inside the tube is turbulent (> 2,300) where its velocity profile is relatively flat. Because the inlet
flow is disturbed through a long gas pipeline and a layer of steel bead, it is expected that the downstream
flow through the tube is quite turbulent. On the other hand, the Reynolds number for the external airflow
over the fuel surface (
) is much smaller than the turbulent limit of , so the
boundary-layer flow on the fuel surface is laminar.
2.3. Test procedure
The biomass sample was ignited by a torch at one end, and then inserted into the middle part of
the wind tunnel and fixed vertically at the tube axis by a sample holder, as shown in Fig. 1(b). The
ignited end (~5 mm) was placed on the bottom for the concurrent smoldering propagation, while for the
opposed propagation, the ignited end was on the top. To reduce the effect of ignition, the smoldering
front was allowed to propagate 20-30 mm away from the ignition region before calculating the
smoldering propagation rate. Afterward, wind with prescribed speed was applied, and shortly after, the
smoldering propagation reached the quasi-steady state (see more details in Fig. A2 in Appendix). The
external wind was applied in a step-increase manner from no wind (i.e., as the base case) until
the critical airflow velocity for blow-off () was found. To start a new test under a different wind
velocity, a fresh fuel sample was used.
A side-view digital video camera was used to capture the time history of the smoldering front.
Through image analysis frame by frame, the instantaneous smoldering propagation rate () can be
calculated as  
, where  is the required duration for a smoldering front to propagate for a
certain distance of . Then, we could judge whether a steady-state propagation was reached (see Fig.
A2 in the Appendix). For each scenario, tests were repeated at least three times to quantify the standard
deviations, and more repeating tests were conducted near the blow-off limit. In general, good
experimental repeatability was found. During the tests, the ambient temperature () was 23 ± 2 oC, and
the relative humidity was 50 ± 10%.
3. Results and Discussion
3.1. Smoldering phenomena
Fig. 2(a) and (b) shows some typical photos of concurrent and opposed smoldering propagation
under different airflow velocities of 0, 5, and 10 m/s with fuel diameters of 1.5, 2.5, and 5.0 mm. As
the wind velocity increased, the smoldering of incense was stronger due to a better oxygen supply,
where the conical reaction surface was hot enough to emit visible light (glowing incandescence) [11].
However, no smoldering-to-flaming transition was observed in this work, different from the past low-
airflow tests [1921]. This was probably because the external wind was already large enough to blow
S. Lin, T. Chow, X. Huang (2021) Smoldering Propagation and Blow-off on Consolidated Fuel under External Airflow,
Combustion and Flame, 234, 111685. https://doi.org/10.1016/j.combustflame.2021.111685
6
off the flame (usually < 5 m/s [4043]). On the other hand, except for oxygen supply, the permeability
of a fuel and its ability to remain consolidated may also affect this transition [33]. Nevertheless,
increasing the oxygen concentration could promote the transition to flame in a smaller airflow [34].
Moreover, compared to the opposed propagation, the glowing zone is brighter for the concurrent
propagation under the same airflow velocity. The length of the glowing zone ( or smoldering front
thickness) increased as the fuel diameter increased, but it was insensitive to the airflow velocity unless
near the blow-off limit. The flat leading edge of the glowing region (not the tip of conical shape) was
used to track the smoldering front. The glowing tip might not be the perfectly conical shape or clearly
observed, because an ash layer sometimes remained and covered the conical tip, just like the burning
cigarette (see the supplementary video). Fig. 2(c) also shows a typical blow-off process for the
smoldering over a 2.5-mm thick incense, where the opposed airflow velocity was increased to 15 m/s.
Gradually, the smoldering (glowing) zone became weaker, flatter, and smaller. After maintaining for
about 3 min, the smoldering was eventually blown off.
Fig. 2 Smoldering propagation on incense rods of 1.5, 2.5, and 5.0-mm diameters under (a) concurrent, and
(b) opposed airflow velocities of 0, 5, and 10 m/s; and (c) blow-off for smoldering on a 2.5-mm incense
under the opposed airflow velocity of 15 m/s.
3.2. Smoldering propagation rate vs. airflow direction
Fig. 3 compares the rate of smoldering propagation at different airflow directions. As expected,
the concurrent smoldering propagation is much faster than the opposed propagation, and the trend of
which is essentially the same as flame spread [11]. For example, for a 2.5-mm thick incense, the
S. Lin, T. Chow, X. Huang (2021) Smoldering Propagation and Blow-off on Consolidated Fuel under External Airflow,
Combustion and Flame, 234, 111685. https://doi.org/10.1016/j.combustflame.2021.111685
7
smoldering propagation rate is around 1.9 cm/min under a concurrent airflow of 5 m/s, tripling that
under an opposed airflow.
In general, the fire spread can be viewed as a continuous ignition process [9,11]. Thus, its rate is
driven by the heat transfer from the oxidation zone () and resisted by the fuel thermal inertia ()
[11] as
 
 


where , ,  , and  are the fuel bulk density, specific heat capacity,
smoldering temperature, and enthalpy change, respectively. The effect of permeability or porosity ()
could be reflected by the difference in bulk density () as , where is the solid
density of biomass sample. For smoldering fire propagation, the preheated length () from glowing
char-oxidation zone to the unburnt zone is close to the thermal penetration depth (, i.e., ,
because both are the characteristic length of heat conduction in solid fuel [11].
As illustrated in Fig. 3(b), for concurrent smoldering propagation, the airflow can directly attack
the conical reaction front, so partial airflow may permeate into the porous glowing zone in the form of
a Darcy flow. The excessive oxygen supply intensifies the char oxidation and increases smoldering
temperature (see intense incandescence in Fig. 2(a)), so a larger preheating flux () will be conducted
from the reaction front to the preheated zone. In addition, the conical glowing zone may preheat the
airflow boundary layer, which can preheat the downstream unburnt fuel via convection. Both effects of
the concurrent airflow can speed up the smoldering propagation.
Fig. 3. (a) Comparison of smoldering propagation rate under external concurrent and opposed airflow,
where the markers show the average values and error bars show the standard deviations, and (b) schematic
diagrams of smoldering propagation under concurrent and opposed airflow.
In contrast, for the smoldering propagation under opposed airflow, the cool airflow can directly
cool the unburnt zone, reducing the preheating from the hot glowing zone () to the preheat zone.
S. Lin, T. Chow, X. Huang (2021) Smoldering Propagation and Blow-off on Consolidated Fuel under External Airflow,
Combustion and Flame, 234, 111685. https://doi.org/10.1016/j.combustflame.2021.111685
8
Furthermore, the oxygen can only reach the char surface via diffusion of the boundary layer, rather than
the pressure-driven Darcy flow under concurrent airflow. Thus, the oxygen supply from the opposed
airflow is less sufficient, slowing down the smoldering propagation. The relatively limited oxygen
supply of opposed smoldering is also reflected by a weaker glowing zone in Fig. 2(b).
3.3. Effect of airflow velocity
Fig. 3 also illustrates the effect of airflow velocity on the smoldering propagation rate, where a
similar trend is found for both concurrent and opposed propagations (see more comparisons in Figs. 4a-
b and 5a-b). That is, as the external airflow velocity increases, the smoldering propagation rate first
increases rapidly to the maximum value (O2-limited Regime) and then remains constant over a wide
range of airflow velocities (Thermal Regime). Subsequently, the propagation rate slightly decreases
(Chemical Regime) until blow-off, following a similar pattern of concurrent flame spread [41,42].
In a small-airflow regime, the smoldering temperature increases with airflow velocity, indicated
by a brighter glowing zone. Therefore, oxygen supply controls the smoldering propagation in this
regime, while the cooling effect of airflow is negligible. For example, as the concurrent airflow velocity
increases from 0 m/s to 3 m/s, the rate of smoldering propagation on the 2.5-mm thick fuel
monotonically increases from 0.8 cm/min to 1.6 cm/min. Such an increasing trend is defined as the O2-
limited Regime, referring to the terminology widely used for the opposed flame spread [11,14,17].
For a consolidated fuel, the smoldering propagation could be regarded as a burning or fuel-
regression process, similar to the burning of a candle or the premixed flame [11,15,29]. Therefore, the
smoldering propagation rate () is the same as the regression rate () as
 




(O2-limited Regime)
where is the velocity of internal airflow inside the conical porous char. Its magnitude could be
estimated by Darcy’s law dominated in the concurrent smoldering and by the diffusion within the
boundary layer dominated in the opposed smoldering as


(concurrent)


(opposed) 
where the Nusselt number changes with flow velocity and diameter as  with
. For opposed smoldering propagation, the internal airflow velocity still changes with the external
airflow () but is several orders of magnitude smaller than that of concurrent smoldering propagation.
Therefore, the smoldering propagation rate at the O2-limited Regime increases with the airflow velocity,
regardless of the flow direction (see Fig. 3a).
S. Lin, T. Chow, X. Huang (2021) Smoldering Propagation and Blow-off on Consolidated Fuel under External Airflow,
Combustion and Flame, 234, 111685. https://doi.org/10.1016/j.combustflame.2021.111685
9
Continuously increasing the airflow velocity, the smoldering propagation rate becomes stable. For
example, the concurrent propagation rate on the 2.5-mm thick fuel remains at 2.1 ± 0.3 cm/min from 7
m/s to 23 m/s in Fig. 3(a), regardless of the airflow velocity. In this large-airflow regime, the unlimited
oxygen supply no longer affects the smoldering propagation rate. Instead, the thermal conduction within
the fuel ( ) starts to dominate the smoldering propagation [11]. This behavior is
similar to the Thermal Regime of the flame spread, where the preheating of flame controls the rate of
flame spread [42,47]. Based on Eq. (1), the smoldering propagation rate at the Thermal Regime is free
of oxygen effect and reach the maximum value as
  
  
 
(Thermal Regime)
where  and  are the fuel thermal conductivity and diffusivity, and is the thermal length within
the fuel. Therefore, the Thermal-Regime smoldering propagation rate is insensitive to the external
airflow velocity.
Fig. 4 Effect of fuel diameter on the rate of smoldering propagation under (a) concurrent and (b) opposed
airflow, (c) maximum smoldering propagation rate and (d) blow-off limits.
Further increasing the external airflow velocity, the smoldering propagation rate eventually starts
to decrease, where the cooling effect of external airflow (see Fig. 3b) on char-oxidation reaction at the
S. Lin, T. Chow, X. Huang (2021) Smoldering Propagation and Blow-off on Consolidated Fuel under External Airflow,
Combustion and Flame, 234, 111685. https://doi.org/10.1016/j.combustflame.2021.111685
10
smoldering front can no longer be neglected. Then, the smoldering propagation rate is controlled by the
competition between smoldering heat release and environmental cooling as
 
 

 


 (Chemical Regime) 
where  and  are the rate and heat of smoldering reaction, respectively. Analogous to the flame
spread [11,17], such a smoldering propagation is called the Chemical Regime [11]. Thus, as the airflow
increases, the convective cooling (
 ) increases to slow down the smoldering propagation. Eventually,
the cooling rate of airflow may equal or exceed the heat release rate of smoldering (
 
 ), so
the blow-off or the quenching by airflow occurs (discussed more in Section 3.4). Similar smoldering
extinction behaviors were also observed in the quenching by the cold wall [29] and fuel moisture.
3.4. Smoldering blow-off limits
Table 1 and Fig. (4d, 5d) summarize the blow-off limits of both concurrent and opposed
smoldering propagation over incenses with different fuel diameters and densities. Clearly, the blow-off
of concurrent smoldering propagation is much more difficult than opposed smoldering propagation. For
example, for 2.5-mm thick incense, the blow-off limits of concurrent and opposed smoldering
propagation are 30 m/s and 14 m/s, respectively. As discussed in Section 3.2 and Fig. 3(b), compared
to the smoldering propagation under concurrent airflow, the opposed airflow can directly attack the
preheated zone, thus increasing cooling efficiency on the unburnt fuel. Therefore, smoldering
propagation is easier to achieve blow-off under opposed airflow. Such a trend is also similar to the
flame spread, where the blow-off of opposed flame spread can be achieved in a smaller wind speed [24].
On the other hand, as shown in Fig. 4(d), when the fuel density is 720 kg/m3, as the fuel diameter
increases from 1.5 mm to 5.0 mm, the blow-off airflow velocity () of smoldering propagation
decreases from 46 m/s to 24 m/s under the concurrent airflow and from 15 m/s and 13 m/s under the
opposed airflow, respectively. Similarly, as shown in Fig. 5(d), the blow-off limits of both concurrent
and opposed smoldering decrease as the fuel density increases from 720 kg/m3 to 1,100 kg/m3 with the
same fuel diameter of 1.5 mm (see more analysis in Section 3.5 and 3.6).
More importantly, all the blow-off airflow velocities of smoldering (13-46 m/s) in the present work
are higher than those of flame spread, for example, the concurrent flame spread over the thin wire (2
m/s) [41] and thin cellulose (~5.5 m/s) [48], or the opposed flame spread over PMMA rod (~3m/s) [24],
thin paper/PMMA sheet (~1 m/s) [42] and thin cellulose (0.4-1 m/s) [43]. The observed blow-off airflow
velocity of incense is also higher than 7 m/s of the opposed smoldering propagation over fiberboard [1].
Approximately, the blow-off airflow velocity of smoldering propagation is about one order of
magnitude larger than that of flame spread, so that smoldering is much more persistent than flaming.
From Eq. (5), the blow-off or the quenching by airflow occurs (i.e.,  ) when the cooling rate
of airflow equals to the heat release rate of smoldering as 
 
 , where the cooling flux at the
S. Lin, T. Chow, X. Huang (2021) Smoldering Propagation and Blow-off on Consolidated Fuel under External Airflow,
Combustion and Flame, 234, 111685. https://doi.org/10.1016/j.combustflame.2021.111685
11
extinction limit could be further expressed as [46]

 

For simplicity, by assuming , we obtain

 
 
 
 
Fig. 5 Effect of fuel density on the rate of smoldering propagation under (a) concurrent and (b) opposed
airflow, (c) maximum smoldering propagation rate, , and (d) blow-off airflow velocity, .
To further evaluate the cooling effect of external flow, a smoldering Damkohler number ()
could be proposed referring to the Da of flame, as the ratio of the flow residence time scale () to the
reaction time scale () [11] as





 

 
Similar concept was also proposed for heterogenous combustion of carbon by Tsuji and Matsui [49].
At the blow-off limit, a critical smoldering Damkohler number can be defined from Eqs. (7,8) as
S. Lin, T. Chow, X. Huang (2021) Smoldering Propagation and Blow-off on Consolidated Fuel under External Airflow,
Combustion and Flame, 234, 111685. https://doi.org/10.1016/j.combustflame.2021.111685
12


 
which is essentially a constant depending on fuel properties and flow conditions, like the conventional
critical Damkohler number for the blow-off limit of flame[9]. Note that as the fuel size and geometry
changes, the flow field ( and ) will change, so that the value of 
will change accordingly.
Table 1. The maximum smoldering propagation rate () and blow-off airflow velocity () over
incenses with different fuel diameters and densities.
Diameter
(mm)
Density
(kg/m3)
Maximum smoldering rate, (cm/min)
Blow-off limit,  (m/s)
Concurrent
opposed
Concurrent
opposed
1.5
720
3.2
0.9
46
15
2.5
720
2.1
0.7
30
14
5.0
720
1.2
0.5
24
13
1.5
920
2.1
0.8
37
11
1.5
1,100
1.3
0.6
18
8
3.5. Effect of fuel diameter
Fig. 4(a-b) further compares the effect of fuel diameter () on smoldering propagation under
external airflow. For both concurrent and opposed smoldering propagations, the propagation rate
increases as the fuel diameter decreases. It is consistent with the trend of flame spread in the literature,
i.e., a faster flame spread for a smaller-diameter fuel [28,50]. For example, under the airflow velocity
of 5 m/s, as the fuel diameter increases from 1.5 mm to 5 mm, the concurrent smoldering propagation
rate decreases from 2.1 cm/min to 1.1 cm/min, and the opposed smoldering propagation rate declines
from 0.9 cm/min to 0.5 cm/min. Clearly, the maximum smoldering propagation rate also decreases with
the fuel diameter, as further compared in Fig. 4(c). From Eqs. (2,3), the internal airflow velocity ()
inside the conical porous char is inversely correlated with fuel diameter (), thus the rate of oxygen
supply decreases as the fuel diameter increases. As a result, the rate of smoldering propagation decreases
with the fuel diameter, agreeing with the experimental results in Fig. 4.
The concept of number (i.e., Spalding mass transfer number) has been widely used to estimate
the flaming burning rate of liquid droplet fuels and solids [5153]. Compared to conventional
gasification mass transfer driven by the flame sheet and heat conduction in the gas phase, the pyrolysis
surface for smoldering is driven by the char-oxidation and heat conduction in the solid phase. Thus, the
same concept can be adopted in describing smoldering burning (or propagation). For a cylindrical rod,
the smoldering propagation is two-dimensional in axial and radial directions (see the top view of control
volume in Fig. 6). Considering the smoldering propagation in the radial direction and the analogy with
S. Lin, T. Chow, X. Huang (2021) Smoldering Propagation and Blow-off on Consolidated Fuel under External Airflow,
Combustion and Flame, 234, 111685. https://doi.org/10.1016/j.combustflame.2021.111685
13
flaming burning of droplet [13,52] or cylindrical rod [53], the burning flux (
) of incense can also be
approximated as

 

where is a fitting correlation, and is a constant for a given fuel. Thus, the smoldering rate in the
axial direction is comparable to the observed smoldering rate in the radial direction as
  



which decreases with the fuel diameter [52], agreeing with the experimental results in Fig. 4(a-c).
Because of the curvature effect, the conductive heat flux concentrates towards a smaller radius. A
similar expression is also derived from Eq. (4), with the diameter as the thermal length () as
 


As seen from Fig. 2, the smoldering front thickness () increases as the fuel diameter increases ().
Fig. 6. Schematic diagram of the 2-D (radial and axial) smoldering propagation on a cylindrical fuel and
the primary heat transfer processes.
On the other hand, as discussed in Section 3.4, the blow-off limit of smoldering was found to
decrease as the fuel diameter increases (Fig. 4d). This trend is opposite to the flame spread, where the
blow-off of a thinner fuel occurs at a smaller airflow velocity and the same critical strain rate (
) [24,41]. Fundamentally, the concept of strain rate can be used for flame because the external
wind can pull and bend the gaseous flame sheet. Nevertheless, the smoldering front in the solid phase
cannot be bent like a flame sheet by the external flow. Therefore, the definition of critical strain rate for
blow-off may not be applicable to smoldering combustion.
To explain the influence of fuel diameter on the smoldering blow-off limit (), a simplified
energy conservation equation is applied to the near-limit reaction zone (see the front view of control
volume in Fig. 6). At the blow-off extinction limit, the smoldering rate is zero; the reaction-zone
S. Lin, T. Chow, X. Huang (2021) Smoldering Propagation and Blow-off on Consolidated Fuel under External Airflow,
Combustion and Flame, 234, 111685. https://doi.org/10.1016/j.combustflame.2021.111685
14
thickness is minimal ( ); and the bottom size is already quenched by the large wind. Then, the
heat generation in the oxidation reaction zone is equal to the convective heat loss due to the airflow (
)
and the conduction to the preheat zone (
 ) as

  

where the convective heat loss from the side for the thin oxidation zone is neglected, and the oxidation
rate from the side has reached a maximum (
 ) and can no longer increase with airflow.
Then, the required convective cooling coefficient () can be derived, which also increases with the
increased airflow velocity and the decreased fuel diameter, as


 


where  
  is a smoldering constant, and is assumed for
simplicity. Thus, by rearranging Eq. (13), the dependence of blow-off airflow velocity with fuel
diameter can be expressed as  

Therefore, as the fuel diameter () increases, the required external airflow velocity to blow off
smoldering fire decreases, agreeing with experimental results in Fig. 4(d). Note that if the fuel diameter
further decreases below 1 mm (i.e., an ultra-thin fuel), the strong wind may easily break and remove
the smoldering zone. Then, the extinction is no longer a blow-off but a fuel-removal, which needs
further experimental verification.
3.6. Effect of fuel density
Fig. 5(a-b) also shows the effect of fuel (bulk) density on the concurrent and opposed smoldering
propagation rate, where the maximum rate of smoldering propagation was further compared in Fig. 5c.
As expected, as the fuel density decreases, the smoldering propagation rate increases, agreeing with the
theoretical analysis of Eqs. (1,2) where the maximum propagation rate is inversely proportional to the
fuel density ( ). For example, as the fuel density increases from 720 to 1,100 kg/m3 under
the wind velocity of 10 m/s, the smoldering propagation rate decreases from 2.1 cm/min to 1.3 cm/min
for the concurrent spread and from 0.9 cm/min to 0.5 cm/min for the opposed spread, respectively.
As the (bulk) fuel density of porous media increases ( ), its porosity and
permeability decrease. Thus, at the blow-off limit, the maximum airflow into the porous fuel (
 )
decreases, which reduces the value of  in Eq. (14). Moreover, the thermal conductivity of fuel
increases with the density (), so that the radial heat conduction from the reaction zone to the
preheat zone also increases. From Eq. (14), we can see the required blow-off airflow velocity ()
decreases as the fuel density increases, agreeing with the experimental results in Fig. 5(d).
S. Lin, T. Chow, X. Huang (2021) Smoldering Propagation and Blow-off on Consolidated Fuel under External Airflow,
Combustion and Flame, 234, 111685. https://doi.org/10.1016/j.combustflame.2021.111685
15
4. Conclusions
In this work, we use experimental approaches to investigate the smoldering propagation and blow-
off over cylindrical incenses under concurrent and opposed external wind up to 50 m/s. There are no
experimental data on the smoldering propagation at large wind speeds over 10 m/s and the blow-off
limits of persistent smoldering fire before this study. For concurrent smoldering propagation, partial
airflow may permeate into the porous glowing zone in the form of a Darcy flow, while the oxygen can
only reach the char surface via diffusion for opposed smoldering propagation. Also, the conical glowing
zone may preheat the concurrent airflow boundary layer to preheat the downstream unburnt fuel, which
further promotes the concurrent smoldering propagation faster than the opposed propagation.
We also found that the smoldering propagation rate is very sensitive to the airflow rate. As external
airflow velocity increases, the smoldering propagation rate first increases (O2-limited Regime), and then
remains stable at its maximum value for a wide range of airflow velocity (Thermal Regime). Afterwards,
it slightly decreases (Chemical Regime) until blow-off. Comparatively, the flame-spread rate increases
with the wind speed due to increased convective heating rather than increased oxygen supply. This is a
significant difference between smoldering and flaming spread, because smoldering combustion is
controlled by both oxygen supply and heat loss.
We report for the first time that the blow-off airflow velocity of smoldering propagation (13~46
m/s) is around one order of magnitude larger than that of flame spread, and it decreases as the fuel
diameter or density increases. Blowing-off concurrent smoldering propagation is also more difficult
than opposed propagation, similar to the blow-off of flame spread. Future numerical simulations are
needed to reveal the underlying physical and chemical process of smoldering propagation and blow-off
under different airflow velocities.
CRediT authorship contribution statement
Shaorun Lin: Investigation, Writing-original draft, Formal analysis, Resources. Tsz Him Chow:
Investigation, Resources. Xinyan Huang: Conceptualization, Supervision, Writing-review & editing,
Funding acquisition.
Declaration of competing interest
The authors declare no conflicts of interest.
Acknowledgments
This research is funded by the National Natural Science Foundation of China (NSFC) No. 51876183
and the Society of Fire Protection Engineers (SFPE) Educational & Scientific Foundation. The authors
thank Dr. Supan Wang (Nanjing Tech Univ.) for helping conduct thermal analysis of the incense
sample. The comments from Dr. Han Yuan (Hong Kong PolyU) are also acknowledged.
S. Lin, T. Chow, X. Huang (2021) Smoldering Propagation and Blow-off on Consolidated Fuel under External Airflow,
Combustion and Flame, 234, 111685. https://doi.org/10.1016/j.combustflame.2021.111685
16
References
[1] Palmer KN. Smouldering combustion in dusts and fibrous materials. Combustion and Flame
1957;1:12954.
[2] Rein G. Smoldering Combustion. SFPE Handbook of Fire Protection Engineering
2014;2014:581603.
[3] Ohlemiller TJ. Modeling of smoldering combustion propagation. Progress in Energy and
Combustion Science 1985;11:277310.
[4] Torero JL, Gerhard JI, Martins MF, Zanoni MAB, Rashwan TL, Brown JK. Processes defining
smouldering combustion: Integrated review and synthesis. Progress in Energy and Combustion
Science 2020;81:100869.
[5] Restuccia F, Ptak N, Rein G. Self-heating behavior and ignition of shale rock. Combustion and
Flame 2017;176:2139.
[6] Koya L, Gard C. Investigating the Attacks on the World Trade Center. First edit. New York:
The Rosen Publishing Group, Inc.; 2018.
[7] Rein G. Smouldering Fires and Natural Fuels. In: Claire M. Belcher, editor. Fire Phenomena in
the Earth System, New York: John Wiley & Sons, Ltd.; 2013, p. 1534.
[8] Song Z, Huang X, Kuenzer C, Zhu H, Jiang J, Pan X, et al. Chimney effect induced by
smoldering fire in a U-shaped porous channel: A governing mechanism of the persistent
underground coal fires. Process Safety and Environmental Protection 2020;136:13647.
[9] Williams FA. Mechanisms of fire spread. Symposium (International) on Combustion
1977;16:128194.
[10] Gollner MJ, Miller CH, Tang W, Singh A V. The effect of flow and geometry on concurrent
flame spread. Fire Safety Journal 2017;91:6878.
[11] Huang X, Gao J. A review of near-limit opposed fire spread. Fire Safety Journal
2021;120:103141.
[12] Fernandez-Pello AC, Hirano T. Controlling mechanisms of flame spread. Combustion Science
and Technology 1983;32:131.
[13] Quintiere JG. Fundamentals of fire phenomena. John Wiley; 2006.
[14] Fernandez-Pello a. C, Ray SR, Glassman I. Downward Flame Spread In an Opposed Forced
Flow. Combustion Science and Technology 1978;19:1930.
[15] Xie Q, Zhang Z, Lin S, Qu Y, Huang X. Smoldering Fire of High-Density Cotton Bale Under
Concurrent Wind. Fire Technology 2020;56:224156.
[16] Huang X, Restuccia F, Gramola M, Rein G. Experimental study of the formation and collapse
of an overhang in the lateral spread of smouldering peat fires. Combustion and Flame
2016;168:393402.
[17] Kumar A, Shih HY, T’ien JS. A comparison of extinction limits and spreading rates in opposed
and concurrent spreading flames over thin solids. Combustion and Flame 2003;132:66777.
[18] Konno Y, Hashimoto N, Fujita O. Downward flame spreading over electric wire under various
oxygen concentrations. Proceedings of the Combustion Institute 2019;37:381724.
[19] Kadowaki O, Suzuki M, Kuwana K, Nakamura Y, Kushida G. Limit conditions of smoldering
spread in counterflow configuration: Extinction and smoldering-to-flaming transition.
Proceedings of the Combustion Institute 2020;000:19.
S. Lin, T. Chow, X. Huang (2021) Smoldering Propagation and Blow-off on Consolidated Fuel under External Airflow,
Combustion and Flame, 234, 111685. https://doi.org/10.1016/j.combustflame.2021.111685
17
[20] Yamazaki T, Matsuoka T, Nakamura Y. Near-extinction behavior of smoldering combustion
under highly vacuumed environment. Proceedings of the Combustion Institute 2019;37:4083
90.
[21] Sato J, Sato K, Hirano T. Fire spread mechanisms along steel cylinders in high pressure oxygen.
Combustion and Flame 1983;51:27987.
[22] Yamazaki T, Matsuoka T, Nakamura Y. Near-extinction behavior of smoldering combustion
under highly vacuumed environment. Proceedings of the Combustion Institute 2019;37:4083
90.
[23] Lin S, Liu Y, Huang X. Climate-induced Arctic-boreal peatland fire and carbon loss in the 21st
century. Science of the Total Environment 2021;796:148924.
[24] Huang X, Link S, Rodriguez A, Thomsen M, Olson S, Ferkul P, et al. Transition from opposed
flame spread to fuel regression and blow off: Effect of flow, atmosphere, and microgravity.
Proceedings of the Combustion Institute 2019;37:411726.
[25] Gollner MJ, Williams F., Rangwala A. Upward flame spread over corrugated cardboard.
Combustion and Flame 2011;158:140412.
[26] Huang X, Rein G. Downward Spread of Smoldering Peat Fire: the Role of Moisture, Density
and Oxygen Supply. International Journal of Wildland Fire 2017;26:90718.
[27] Zhou Y, Gong J, Jiang L, Chen C. Orientation effect on upward flame propagation over rigid
polyurethane foam. International Journal of Thermal Sciences 2018;132:8695.
[28] Jiang L, He JJ, Sun JH. Sample width and thickness effects on upward flame spread over PMMA
surface. Journal of Hazardous Materials 2018;342:11420.
[29] Lin S, Huang X. Quenching of smoldering: Effect of wall cooling on extinction. Proceedings of
the Combustion Institute 2021;38:501522.
[30] Fernandez-Pello AC. Flame Spread Modeling. Combustion Science and Technology
1984;39:11934.
[31] Torero JL, Fernandez-Pello AC, KitanO M. Opposed forced flow smoldering of polyurethane
foam. Combustion Science and Technology 1993;91:95117.
[32] Kuwana K, Suzuki K, Tada Y, Kushida G. Effective Lewis number of smoldering spread over
a thin solid in a narrow channel. Proceedings of the Combustion Institute 2017;36:320310.
[33] Valdivieso JP, Rivera J de D. Effect of Wind on Smoldering Combustion Limits of Moist Pine
Needle Beds. Fire Technology 2014;50:1589605.
[34] Santoso MA, Christensen EG, Yang J, Rein G. Review of the Transition From Smouldering to
Flaming Combustion in Wildfires. Frontiers in Mechanical Engineering 2019;5.
[35] He M, Ding L, Yu L, Ji J. Effect of density on the smoldering characteristics of cotton bales
ignited internally. Proceedings of the Combustion Institute 2020;000:19.
[36] Wang S, Huang X, Chen H, Liu N. Interaction between flaming and smouldering in hot-particle
ignition of forest fuels and effects of moisture and wind. International Journal of Wildland Fire
2017;26:7181.
[37] Bar-Ilan A, Putzeys OM, Rein G, Fernandez-Pello AC, Urban DL. Transition from forward
smoldering to flaming in small polyurethane foam samples. Proceedings of the Combustion
Institute 2005;30:2295302.
[38] Emberley R, Inghelbrecht A, Yu Z, Torero JL. Self-extinction of timber. Proceedings of the
S. Lin, T. Chow, X. Huang (2021) Smoldering Propagation and Blow-off on Consolidated Fuel under External Airflow,
Combustion and Flame, 234, 111685. https://doi.org/10.1016/j.combustflame.2021.111685
18
Combustion Institute 2017;36:305562.
[39] Dosanjh SS, Pagni PJ, Fernandez-Pello AC. Forced cocurrent smoldering combustion.
Combustion and Flame 1987;68:13142.
[40] Loh HT, Fernandez-Pello AC. Flow Assisted Flame Spread Over Thermally Thin Fuels. Western
States Section, Combustion Institute (Paper) 1984:6574.
[41] Lu Y, Huang X, Hu L, Fernandez-Pello C. Concurrent Flame Spread and Blow-Off Over
Horizontal Thin Electrical Wires. Fire Technology 2019;55:193209.
[42] Fernandez-Pello AC, Ray SR, Glassman I. Flame spread in an opposed forced flow: the effect
of ambient oxygen concentration. Symposium (International) on Combustion 1981;18:57989.
[43] Carmignani L, Celniker G, Bhattacharjee S. The Effect of Boundary Layer on Blow-Off
Extinction in Opposed-Flow Flame Spread over Thin Cellulose: Experiments and a Simplified
Analysis. Fire Technology 2017;53:96782.
[44] Leach S V, Ellzey JL, Ezekoye OA. Convection , Pyrolysis , and Damko Extinction of Reverse
Smoldering Combustion. Twenty-Seventh Symposium (International) on Combustion
1998:287380.
[45] Yamazaki T, Matsuoka T, Li Y, Nakamura Y. Applicability of a Low-Pressure Environment to
Investigate Smoldering Behavior Under Microgravity. Fire Technology 2020.
[46] Incropera FP. Principles of heat and mass transfer. John Wiley; 2007.
[47] De Ris JN. Spread of a laminar diffusion flame. Symposium (International) on Combustion
1969;12:24152.
[48] Wang S, Wang S, Zhu K, Xiao Y, Lu Z. Near Quenching Limit Instabilities of Concurrent Flame
Spread over Thin Solid Fuel. Combustion Science and Technology 2016;188:45171.
[49] Tsuji H, Matsui K. An aerothermochemical analysis of combustion of carbon in the stagnation
flow. Combustion and Flame 1976;26:28397.
[50] Ayani MB, Esfahani JA, Mehrabian R. Downward flame spread over PMMA sheets in quiescent
air: Experimental and theoretical studies. Fire Safety Journal 2006;41:1649.
[51] Rich D, Lautenberger C, Torero JL, Quintiere JG, Fernandez-Pello C. Mass flux of combustible
solids at piloted ignition. Proceedings of the Combustion Institute 2007;31 II:265360.
[52] Kanury AM. Rate of charring combustion in a fire. Symposium (International) on Combustion
1973;14:113142.
[53] Lee CK. Burning rate of fuel cylinders. Combustion and Flame 1978;32:2716.
S. Lin, T. Chow, X. Huang (2021) Smoldering Propagation and Blow-off on Consolidated Fuel under External Airflow,
Combustion and Flame, 234, 111685. https://doi.org/10.1016/j.combustflame.2021.111685
19
Appendix
The incense sample was firstly pulverized into powders for TGA-DSC tests. The initial mass of
peat was about 5 mg, and samples were heated at the constant rates of 10 K/min. Two oxygen
concentrations were selected, 0% (nitrogen) and 21% (air). Experiments were repeated twice for each
case, and good repeatability is shown. Fig. A1 shows the mass-loss rate (DTG) and heat flow (DSC)
curves, respectively. Regardless of the oxygen concentration, the mass-loss rate rapidly increases at
around 250°C, which can be defined as the pyrolysis temperature. The heat of smoldering () can
be calculated by integrating the heat flow curve, and it is about 18 MJ/kg for this incense.
Fig. A1 TGA-DSC results of incense sample under air and nitrogen flow at a heating rate of 10
K/min, (a) normalized mass loss rate; and (b) heat flow as a function of temperature.
Fig. A2 shows some examples of required duration () for a reaction front to propagate through a
certain distance of  under different airflow directions and velocities. Good linearity between  and
indicates the steady-state of smoldering fire propagation, where the slopes of the fitting lines are the
corresponding smoldering propagation rates ().
Fig. A2. Examples of smoldering front position () vs. the experimental duration () over the incense
with different diameters under different airflow velocities.
... In the literature, external oxidizer flow and heat flux are often quantities of interest as they are the key parameters affecting the thresholds of StF transition [34]. For example, airflow or wind is crucial to the StF transition, because it increases both oxygen supply and heat losses [25,35]. An increasing airflow rate increases oxygen supply to the smoldering reaction, leading to a more intense smoldering rate with more pyrolyzate and higher heat release and temperatures that support the occurrence of StF transition. ...
... As wind velocity increases, the smoldering intensity is controlled by a competition between smoldering heat release and environmental cooling [35], where both q ″ sm and q ″ ∞ increases with U. However, as observed from Fig. 3, the required external heat flux increases as the wind velocity increases, which indicates that the StF process is more sensitive to environmental cooling than oxygen supply by the oxidizer flow. ...
... It is difficult to judge the glowing char cracks' effect on the heat transfer and combustion of solid wood. Almost all cellulose fuels can sustain both forms of flaming and smouldering combustion [11,12], the smouldering process can lead to flaming combustion normally in powdery or porous solid fuels [13][14][15], which mainly contain polyurethane foams [16,17], cotton [18], or discrete cellulose powders like particle board, sawdust and shredded wood [14]. Smouldering-to-flaming (StF) transition is a complex process in which the smoulder reaction provides the heat to pyrolyze the virgin fuel, or acts as a pilot to ignite the flammable gaseous mixture, involving the transition from "heterogeneous reactions" to "homogeneous gas phase reactions". ...
... Eqs. (11)(12)(13)(14)(15)(16)(17) are solved by finite element methods, using the GMRES (Generalized Minimal ...
Article
Full-text available
The shrinkage, deformation and cracking of the wood affect their smouldering and flaming dynamics, but the scientific understanding is still limited. We study the burning behaviours of disc wood samples with a diameter of 60 mm and thicknesses of 5-40 mm under external airflows up to 6 m/s. Results show that the smouldering-to-flaming (StF) transition can be observed at about 830 • C under external airflow, which is caused by the interactions between smouldering-induced crack and environmental airflow. The fully penetrated vertical char crack or pre-perforated hole promotes the StF transition because of (1) enhanced radiation between the two smouldering surfaces and (2) greater air supply under the chimney effect. As the wind velocity increases, both the smouldering surface temperature and crack size increase, so the transition to flaming becomes faster. For a larger wood thickness, a larger airflow is required to generate the crack and cause a StF transition. A numerical model is proposed to investigate the volatile convection and flaming ignition. Numerical analysis reproduces the StF transition, as an autoignition of a pyrolysate-oxygen mixture promoted by hot smouldering surfaces. The numerical model further reveals the effects of smouldering temperature and cross wind on the StF transition. This work deepens the understanding of the StF transition dynamics and provides insights into the wildfire ignition dynamics and fire hazards of timber structures.
... Fire is a key disturbance to peatlands globally, driven primarily by smouldering, leading to one of the most persistent types of fire phenomena on Earth (Rein 2013;Lin et al. 2021a). Smouldering is a heterogeneous reaction sustained by exothermic oxidation in the solid phase, often characterised as a sluggish, flameless, and low-temperature combustion process (Ohlemiller 1986;Lin et al. 2021c). Although peatlands only cover 2-3% of the global land surface, they store over 25% of terrestrial organic carbon, which is approximately equal to that stored in the atmosphere or living plants (Lehmann and Joseph 2015;Kohlenberg et al. 2018). ...
Article
Full-text available
Background Wildfires represent a significant threat to peatlands globally, but whether peat fires can be initiated by a lofted firebrand is still unknown. Aims We investigated the ignition threshold of peat fires by a glowing firebrand through laboratory-scale experiments. Methods The oven-dried weight (ODW) moisture content (MC) of peat samples varied from 5% ODW to 100% ODW, and external wind (ν) with velocities up to 1 m/s was provided in a wind tunnel. Key results and conclusions When MC < 35%, ignition is always achieved, regardless of wind velocity. However, if MC is between 35 and 85%, an external wind (increasing with peat moisture) is required to increase the reaction rate of the firebrand and thus heating to the peat sample. Further increasing the MC to be higher than 85%, no ignition could be achieved by a single laboratory firebrand. Finally, derived from the experimental results, a 90% ignition probability curve was produced by a logistic regression model. Implications This work indicates the importance of maintaining a high moisture content of peat to prevent ignition by firebrands and helps us better understand the progression of large peat fires.
... For this geometry, wind speeds on the order of 1 m/s were beneficial for the smoldering process (measured in terms of mass loss and CO production rates), but fuel porosity did not seem to have a large impact on the results due to the large void fraction of the cribs. Porosity and air flow had higher impact on woody fuel beds such as pine bark [25], peat [26,27], and incense [28]; these studies also showed the importance of sample size and geometry on the heat losses to the surrounding environment and thus the sustainability of the smoldering process. The forced air flow through the porous fuels was beneficial at low and intermediate velocities, in analogy with other non-woody fuels such as polyurethane foam [29,30], but high velocities eventually led to blow-off extinction. ...
Article
Large and downed woody fuels remaining behind a wildfire's flame front tend to burn in a smoldering regime, producing large quantities of toxic gases and particulate emissions, which deteriorates air quality and compromises human health. Smoldering burning rates are affected by fuel type and size, the amount of oxygen reaching the surface, and heat losses to the surroundings. An external wind has the dual effects of bringing fresh oxidizer to the fuel surface and porous interior, while at the same time enhancing convective cooling. In this work, a series of experiments were conducted on single and adjacent poplar dowels to investigate the effect of fuel geometry and wind speed on smoldering of woody fuels, including its burning rate and combustion products. Dowels had variable thickness (19.1 and 25.4 mm), aspect ratios, and arrangement (number of dowels and spacing between them). Using measurement of mass loss, CO, and HC production as indicators of the smoldering intensity , the results indicate that the arrangement of smoldering objects significantly affects burning rates and emissions. Specifically, spacings of 1/8 and 1/4 of the dowel thickness enhanced the smoldering process. The smoldering intensity was also enhanced by increased external wind (ranging between 0.3 m/s and 1.5 m/s), but its effect was dependent upon the spacing between the dowels. The convective losses associated with the spacing were further investigated with a simplified computational model. The simulations show that the wind significantly increases convective losses from the smoldering surfaces, which in turn may offset the increase in smoldering intensity related to the higher oxygen flux at higher wind speeds.
... The propagation and extinction of smouldering combustion are primarily controlled by two key mechanisms: heat loss and oxygen supply [2,3,22,23]. The path and rate of oxygen supply are crucial for the heterogeneous oxidations that generate the required heat to balance endothermic processes such as drying, pre-heating, environmental cooling, and endothermic pyrolysis reactions during propagation [21]. ...
Article
Full-text available
Smouldering is a low-temperature, flameless, and persistent combustion process driven by heterogeneous oxidations. Oxygen supply is a key parameter of smouldering and is sensitive to fuel density and particle size, but our understanding is still limited. Herein, we explore the oxygen threshold for smouldering propagation under upward internal airflow velocities up to 5 mm/s. Pine needles with different bulk densities (55-120 kg/m 3) and wood samples with different particle sizes (1-50 mm) are tested. We found that the minimum airflow velocity for sustaining smouldering propagation increases with the decrease of the bulk density or the increase of the particle size. By increasing the airflow velocity, the smouldering front first propagates unidirectionally (opposed) and then bidirectionally (opposed + forward). Nevertheless, when the pore size is large (the fuel particle size is large or the fuel bulk density is small), bidirectional propagation always occurs, because the oxygen can leak through the opposed smouldering front. A simplified thermochemical analysis is proposed to reveal the influence of interparticle heat transfer on the minimum oxygen supply rate of smouldering propagation. This work advances the fundamental understanding of smouldering on solid fuel particles and their smouldering fire risks and helps optimize the efficiency and safety of smouldering processes.
... The flame blow-off has been widely studied in lab experiments under well-controlled external airflows or winds. Two mechanisms for flame blow-off are proposed, (1) within the flame sheet, the reduced flow residence time scale compared to the reaction time scale required for a continuous burning [10], and (2) in case of fire, a strong wind can also remove the solid fuel directly or reduce the gaseous fuel release from the solids by cooling [11], breaking the flame-fuel-supply cycle and leads to extinction [9,12]. Depending on the fuel type and flame characteristics, some critical blow-off velocity [13,14], stretch rate [15,16], or Damköhler number [10,[17][18][19] can be used to quantify the flame blow-off limit. ...
Article
Full-text available
The blow-off is one of the typical flame extinction mechanisms, and such a principle has been widely used in firefighting when the water-based extinguisher is limited. This work explores the blow-off extinctions of different diffusion flames by air vortex ring. The vortex ring harnesses its kinetic energy within the fast-rotating vortex core, enabling the transmission of power over several meters to blow off a remote flame. The power required for vortex ring blow-off is found to be two to three orders of magnitude smaller than the power of the flame itself, demonstrating exceptional energy efficiency. It is observed that the poloidal flow (circulation) surrounding the vortex core can stretch the flame base to the critical state and then cause instantaneous extinction. To explain the vortex-induced blow-off limit, a critical Damköhler number that accounts for the competition between fuel gas flow and flame stretch was formulated. This work provides a fundamental understanding of the extinction mechanism by vortex ring, and it offers technical guidelines for using air as a flame extinguisher for remote firefighting within minimum energy input.
Article
Full-text available
Smoldering treatment is emerging as a valuable engineering tool for many processes, including food waste treatment. However, smoldering systems are currently not well-understood nor optimized. Therefore, numerical models provide invaluable insight into the process dynamics, which improves our understanding and supports the development of novel systems. These smoldering models couple heat, mass, and momentum transfer with pyrolysis and oxidation chemical reactions within porous media. While recent models have untangled many aspects of these systems, local oxygen transport rates from bulk flow to the fuel surface are still not well-resolved. In this work, local oxygen-transport equation was approximated by an analytic derivation based on the gas–solid oxygen non-equilibrium hypothesis. With the improved oxygen-transport equation, a 2D model with five-step reaction scheme for smoldering propagation of food waste in sand was developed. Kinetic parameters obtained from TG experiments were incorporated into the bed-scale smoldering propagation model. The developed model was validated with experimental data that stretched from robust to weak smoldering propagation. It was demonstrated that the developed model matches well with experiments. Furthermore, this model revealed: (i) the emergence of non-uniform gas flow in the reactor, (ii) the evolution of the kinetic- and oxygen-transport-limiting regimes, and (iii) valuable insight into the fundamental changes with smoldering robustness.
Article
Full-text available
Boreal peatlands are increasingly vulnerable to wildfires as climate change continues accelerating. Fires consume substantial quantities of organic soils and rapidly transfer large stocks of terrestrial carbon to the atmosphere. Herein, we quantify the minimum environmental temperature that allows the moist peat to smolder, as the fire threshold of peatlands. We then apply a typical vertical soil temperature profile to estimate the future depth of burn and carbon emissions from boreal peatland fires under the impact of global warming. If the boreal region continues warming at a rate of 0.44 °C/decade, we estimate the carbon loss from the boreal peat fires on a warmer soil layer may increase from 143 Mt. in 2015 to 544 Mt. in 2100 and reach a total of 28 Gt in the 21st century. If the global human efforts successfully reduce the boreal warming rate to 0.3 °C/decade, the peat fire carbon loss would drop by 21% to 22 Gt in the 21st century. This work helps understand the vulnerability of boreal peatland to more frequent and severer wildfires driven by global warming and estimate climate-induced carbon emissions from boreal peatland fires in the 21st century.
Article
Full-text available
The flame spread over combustible materials is often affected by the fire thermal radiation and convection and the heat exchange with adjacent objects, which are especially complex on melting thermoplastics. This work chooses polyethylene (PE) tubes with a 2-mm thin wall to study the flame-spread behaviors under three heating conditions, (a) hot inner boundary, (b) hot ambient, and (c) external radiation. The tubes could simulate the insulation of electrical wires, and the inner boundary was controlled by flowing oil through at a constant temperature. Results show that just above the fuel molten point, the flame-spread rate unexpectedly decreases with the increasing environmental temperature, because the conductive cooling changed to convective cooling of molten PE. A thin layer of fuel can remain after the flame spread, and as the boundary temperature increases, the remaining PE decreases while the dripping mass increases. Under intense heating, burning behaviors eventually become similar regardless of the heating scenario. This work helps understand the flame spread and phase change of thermoplastic fires, particularly wires and cables, under various heating scenarios of realistic fire events.
Article
Full-text available
Smouldering wildfire is an important disturbance to peatlands worldwide, and it contributes significantly to global carbon emissions and provides positive feedback to climate change. Herein, we explore the feasibility of firebreaks to control smouldering peat fires through laboratory-scale experiments. The dry-mass moisture content (MC) of peat soil varied from 10% (air-dried) to 125%. We found that smouldering peat fire may be successfully extinguished above the mineral soil layer, even if the peat layer is not entirely removed. There are two criteria for an effective peat firebreak: (I) adding water to make the peat layer sufficiently wet (> 115% MC in the present work), and (II) ensuring that the peat layer is thinner than the quenching thickness (< 5 cm). Criterion I may fail if the water table declines or the peat layer is dried by surface fires and hot weather, thus satisfying Criterion II is more attainable. A sloped, trench-shaped firebreak is recommended to guide water flow and help maintain high peat moisture content. This work provides a scientific foundation for fighting and mitigating smouldering wildfires and guides protective measures for field-scale peat fire experiments.
Article
Full-text available
Smoldering is the slow, low-temperature, and flameless combustion phenomenon in porous fuels. Smoldering is different from flaming regarding the chemical and transport processes, despite sharing many similarities in ignition and fire spread. In this work, we explore the applicability of quenching and quenching diameter in smoldering combustion. The smoldering of dry organic soil is initiated in the 25-cm long tubular reactor with different diameters from 4 cm to 15 cm. The thermal boundary and oxygen supply of the smoldering reactor are varied by using different wall materials and opening configurations, respectively. The quenching of smoldering is observed as the diameter of the reactor decreased, which is the same as the quenching of the premixed flame. The minimum smoldering temperature (∼250 °C) and propagation rate (∼0.5 cm/h or 0.1 mm/min) are found before quenching. The measured quenching diameter of smoldering is about 10 cm (much larger than the flame) and comparable to the thickness of reaction front (similar to the flame). The quenching diameter of smoldering increases as the wall cooling increases and the oxygen supply decreases. The influence of oxygen supply is unique to the smoldering quenching phenomenon because it affects the mode of smoldering propagation. This work helps understand the persistence and extinction limit of smoldering and supports the prevention and suppression strategies for smoldering fire.
Article
Full-text available
Creeping fire spread under opposed airflow is a classic fundamental fire research problem involving heat transfer, fluid dynamics, chemical kinetics, and is strongly dependent on environmental factors. Persistent research over the last 50 years has established a solid framework for different fire-spread processes, but new fire phenomena and recent developments continue to challenge our current understanding and inspire future research areas. In this review, we revisit the problem of opposed fire spread under limited and excessive oxygen supply. Various near-limit fire phenomena, as recently observed in flaming, smoldering, and glowing spread under various environment and fuel configurations, are reviewed in detail. Particularly, aspects of apparent importance, such as transition phenomena and heterogenous chemistry, in near-limit fire spread are highlighted, and valuable problems for future research are suggested.
Article
Full-text available
Cotton is the most widely used natural textile fiber for human beings, and fire safety during transportation, storage, and manufacturing is of great significance. This work investigated the smoldering burning of a high-density cotton bale (36 L and 225 kg/m3) tightened by thin steel wire ropes under the concurrent wind. Without wind, the creeping smoldering spread showed two stages: (I) the relatively fast surface spread until the smoldering front covered the entire free surface (13 cm/h), and then (II) the slow in-depth spread from the surface to the internal (4 cm/h). With a concurrent wind, only one major concurrent smoldering front was observed from the free surface to the internal fuel, where the rate of smoldering spread decreased as the sample depth increased. The smoldering spread rate and peak temperature were found to increase almost linearly with the wind velocity due to the enhanced oxygen supply. The steel wire rope could appreciably obstruct the free-surface spread and slow down the overall smoldering spread. A large wind occasionally led to a smoldering-to-flaming transition, but the flame could not be sustained. This research improves our understanding of the wind effect on the smoldering spread in the compressed porous biomass and helps evaluate the fire risk of cotton during transportation and storage.
Article
The cotton fire often occurs during storage, and it is valuable to reveal the combustion characteristics of cotton bales. However, most of the previous studies only focused on the combustion phenomenon of cotton bales after ignited externally. In fact, during the storage of cotton, the fire often originated from inside of a cotton bale, resulted from the sparks introduced into the cotton bale by the ginning machinery and the heat generated due to hygroscopic property. Therefore, an experimental study was conducted to investigate the effect of density on the smoldering characteristics of cotton bales ignited internally. In this study, the cotton densities of experimental study were 50, 60, 70, 80, 90, 100, 110 and 120 kg/m³. The change rules of the smoldering spread rate in external horizontal, internal horizontal, and internal vertical during the combustion process as well as the mass loss of the cotton bales with these densities were studied, and the spreading processes of the smoldering in these cotton bales were also determined. The experimental results show that due to differences in the amount of oxygen and heat flux, an increase in density leads to a decrease in the spread rates in external horizontal, internal horizontal and internal vertical. With the increase in density, the porosity decreases and the amount of air is reduced, so the spread rate gradually decreases. However, the mass loss rate, i.e. the amount of burning cotton per unit time, increases with density. And, as the overall amount of cotton increases with density, the burning time also increases.
Article
This paper presents an experimental and theoretical study on limit conditions for smoldering combustion of biomass fuel. A fuel rod was burned in an oxidizer flow in a Tsuji-Yamaoka type, counterflow configuration, and the spread rate and temperature distribution were measured. The spread rate and maximum temperature increased with an increase in oxygen mass fraction in the oxidizer flow. Smoldering spread was not possible when the oxygen mass fraction was below a lower critical value, whereas a smoldering-to-flaming transition (SFT) occurred when the mass fraction was above an upper critical value. A one-dimensional model was then developed based on experimental observations to describe steady smoldering spread. The limit conditions were predicted by assessing the existence of solution. It was found that extinction occurs due to finiteness of char-oxidation rate, while the SFT limit can be predicted by considering gas-phase combustion of pyrolysis products.
Article
Smouldering combustion is an important and complex phenomenon that is central to a wide range of problems (hazards) and solutions (applications). A rich history of research in the context of fire safety has yet to be integrated with the more recent, rapidly growing body of work in engineered smouldering solutions. The variety of disciplines, materials involved, and perspectives on smouldering have resulted in a lack of unity in the expression of key concepts, terminology used, interpretation of results, and conclusions extracted. This review brings together theoretical, experimental, and modelling studies across both fire safety and applied smouldering research to produce a unified conceptual understanding of smouldering combustion. The review includes (i) an overview of the fundamental processes with a synthesis of nomenclature to generate a consistent set of terms for these fundamental processes, (ii) a distillation of ignition, extinction, and transition to flaming research, (iii) a review of the temporal and spatial distribution of heat and mass transfer processes as well as their solution using analytical and numerical methods, (iv) an overview of smouldering emissions and emission treatment systems, and (v) a summary of key gaps and opportunities for future research. Beyond merely review, a new conceptual model is provided that articulates similarities and critical differences between the two main smouldering systems: porous solid fuels and condensed fuels in inert porous media. A quantitative analysis of this conceptual model reveals that the evolution of a smouldering front, while a local process, is determined by a global energy balance that is cumulative in time and has to be integrated in space. As such, the fate of a smouldering reaction can be predicted before the effects of global heat exchange have impacted the reaction. This approach is relevant to all forms of smouldering (including fire safety), but it is particularly important when using smouldering as an engineered process that results in the positive use of the energy released by the smouldering reaction (applied smouldering). In applied smouldering, predicting the fate of a reaction ahead of time allows operators to modify the conditions of the process to maintain self-sustained smouldering propagation and thus fully harness the benefits of the reaction.