You are on page 1of 14

Combustion Science and Technology

ISSN: 0010-2202 (Print) 1563-521X (Online) Journal homepage: http://www.tandfonline.com/loi/gcst20

Wrinkling of Large-Scale Flame in Lean


Propane–Air Mixture Due to Cellular Instabilities

Wookyung Kim, Takuma Endo, Toshio Mogi, Kazunori Kuwana & Ritsu
Dobashi

To cite this article: Wookyung Kim, Takuma Endo, Toshio Mogi, Kazunori Kuwana & Ritsu
Dobashi (2018): Wrinkling of Large-Scale Flame in Lean Propane–Air Mixture Due to Cellular
Instabilities, Combustion Science and Technology

To link to this article: https://doi.org/10.1080/00102202.2018.1502757

Published online: 06 Aug 2018.

Submit your article to this journal

View Crossmark data

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=gcst20
COMBUSTION SCIENCE AND TECHNOLOGY
https://doi.org/10.1080/00102202.2018.1502757

Wrinkling of Large-Scale Flame in Lean Propane–Air Mixture


Due to Cellular Instabilities
Wookyung Kim a, Takuma Endoa, Toshio Mogib, Kazunori Kuwanac,
and Ritsu Dobashid
a
Department of Mechanical Systems Engineering, Hiroshima University, Higashihiroshima, Japan; bGraduate
School of Engineering, The University of Tokyo, Tokyo, Japan; cDepartment of Chemistry and Chemical
Engineering, Yamagata University, Yamagata, Japan; dDepartment of Chemical System Engineering, The
University of Tokyo, Tokyo, Japan

ABSTRACT ARTICLE HISTORY


In accidental gas explosions, flame acceleration owing to cellular Received 13 March 2018
instabilities such as diffusional–thermal instability and Darrieus– Revised 5 July 2018
Landau instability can cause considerable damages, for example, Accepted 17 July 2018
the formation of a strong blast wave. In particular, as the flame KEY WORDS
scale increases, Darrieus–Landau instability, caused by a density Gas Explosion; Flame
jump, progressively dominates in the flame acceleration. In this Acceleration; Cellular
study, we experimentally investigated the growth and wrinkling Instability; Darrieus–Landau
owing to Darrieus–Landau instability of a spherically expanding Instability; Fractal Structure
flame in a large-scale experiment, in which a propane-air mixture of
the equivalence ratio ϕ = 0.8 was filled and ignited in a plastic tent of
27 m3. Experimental images of large-scale flames of the lean pro-
pane–air mixture, for which the flame is diffusional–thermally stable,
were analyzed. The edge of flame was detected and rearranged in
polar coordinates. The results show that small-scale cells merge and
form a bigger cell. The generated bigger cell grows by the instability
mechanism and eventually forms a large single cusp. In addition, the
peak-to-peak amplitude of the wrinkled flame was evaluated. The
value of peak-to-peak amplitude increased as time progressed. Such
a cellular flame gives rise to a fractal-like structure and acceleration of
its propagation speed. The fractal dimension of the wrinkled flame
surface was evaluated by logarithmically plotting the flame speed
versus its radius and also by a box-counting method. The results
demonstrated that the wrinkled structure of a large-scale flame can
be characterized by its fractal dimension and that a transition period
into a well-developed self-similar regime exists.

Introduction
Understanding flame-propagation behavior is of great importance to assess the conse-
quence of an accidental gas explosion because of its significant influence on the blast-wave
intensity (Kim et al., 2015b; Dobashi et al., 2011). It is known that the intensity of a blast
wave is dramatically increased by flame acceleration due to the presence of obstacles and/
or cellular instability mechanisms such as diffusional–thermal and Darrieus–Landau
instabilities (Kim et al., 2015b). When the Lewis number, Le, is less than unity, a flame
is diffusional–thermally unstable, whereas it is stable for Le > 1. On the other hand, every

CONTACT Wookyung Kim youwoo2@gmail.com


Color versions of one or more of the figures in the article can be found online at www.tandfonline.com/gcst.
© 2018 Taylor & Francis
2 W. KIM ET AL.

flame is unstable in terms of Darrieus–Landau instability that is caused by the density


jump across a flame (Darrieus, 1938; Landau, 1944; Williams, 1985). Darrieus–Landau
instability takes effect when the dimensionless flame radius, the ratio of flame radius to
flame thickness, exceeds a certain critical value (Okafor et al., 2016).
To illustrate the effects of these instability mechanisms on flame shape, hydrogen-air
flames observed in small-scale experiments (the chamber volume, V = 0.79 L) (Kim et al.,
2018) are shown in Figure 1. The rich hydrogen flame (Le > 1) at the atmospheric pressure
in Figure 1(a) was stable with few wrinkles formed on the surface, while the lean hydrogen
flame (Le < 1) at the atmospheric pressure in Figure 1(b) wrinkled owing to diffusional–
thermal instability. The rich hydrogen flame at 0.5 MPa in Figure 1(c) was wrinkled owing
to Darrieus–Landau instability enhanced by the increased dimensionless flame radius due
to a decreased flame thickness at the elevated pressure, although the flame is diffusional–
thermally stable. The lean hydrogen flame at 0.5 MPa in Figure 1(d) was wrinkled by
diffusional–thermal instability as well as Darrieus–Landau instability. These results
demonstrate that the wrinkled structures of flames depend on the type of instability.
The wrinkled flame structures due to diffusional–thermal and Darrieus–Landau instabil-
ities have been investigated numerically by Gutman et al. (Gutman and Sivashinsky, 1990)
and Mukaiyama et al. (Mukaiyama et al., 2013). Their numerical results showed that flame
corrugations formed by an initial disturbance of short wavelengths merge and coalesce
into large single cusps owing to Darrieus–Landau instability. In addition, their results

Figure 1. Observation of spherically expanding hydrogen-air flames for various pressures and equiva-
lence ratios: (a) stable flame, (b) unstable flame due to diffusional–thermal instability, (c) unstable flame
due to Darrieus–Landau instability, and (d) unstable flame due to Darrieus–Landau and diffusional–
thermal instabilities (Kim et al., 2018).
COMBUSTION SCIENCE AND TECHNOLOGY 3

illustrated that the flame structure due to Darrieus–Landau instability has fractal char-
acters, whereas the structure formed by diffusional–thermal instability does not. As the
flame size increases, Darrieus–Landau instability manifests and gradually dominates its
accelerative motion. Therefore, the intensity of blast wave in a large-scale accidental gas
explosion is mainly affected by flame acceleration owing to Darrieus–Landau instability.
Although Darrieus–Landau instability is a main factor influencing flame acceleration,
there are few experimental studies focusing on the detailed cellular structures of large-
scale flames influenced by Darrieus–Landau instability (Clanet and Searby, 1998).
In view of the above consideration, we investigated the growth and wrinkling of a
spherically expanding large-scale flame which is diffusional–thermally stable but wrinkled
by Darrieus–Landau instability. This paper analyzes the edge of flame front and discusses
its structure. In addition, in order to apply the prediction model to large-scale gas
explosions, the fractal dimension of wrinkled flame surface was evaluated by logarithmi-
cally plotting the propagation speed versus the flame radius. Moreover, in order to
investigate the self-similarity of large-scale flames, the fractal dimension was calculated
also by a box-counting method.

Experimental specifications
Field experiments were conducted using a plastic tent consisting of a cubic stainless-steel
frame covered with a polyethylene plastic film of 0.1 mm thick. The volume of the tent is
V = 27 m3. A propane-air mixture confined in the tent was ignited at its center, and flame
images propagating through the mixture were recorded using high speed cameras at 2000
frames per second. The blast wave generated by gas explosion at various distances from
the center was simultaneously measured by piezoelectric pressure sensors. The experi-
mental procedure is given in more detail in Ref. (Kim et al., 2015a). In this paper, we
exclusively discuss a large-scale propane-air flame of the equivalence ratio ϕ = 0.8, whose
Lewis number is greater that unity, in order to investigate the sole effects of Darrieus–
Landau instability on the growth and wrinkling of the flame.
In this study, we focus on the structure of a large-scale flame captured by one of the
field experiments. The edge of the flame front was detected by the Canny edge-detection
method, which is used in MATLAB. Figure 2 shows an original and the corresponding
edge-detected images of the large-scale propane-air flame of ϕ = 0.8, for which the flame is
diffusional–thermally stable, at a flame radius of about r = 1 m. The evolution of flame
front is shown in Figure 3. The detected images do not contain the formation of small
cells, because of the limitation of high-speed image resolution (2048 × 2048 pixels for
6.3 m × 6.3 m). The resolution is nevertheless good enough to capture grown cells. The
detected flame edge was then rearranged in polar coordinates as shown in Figure 4. The
angle of 90°Corresponds to the upward direction, while 270° to the downward direction.

Results and discussion


Growth of cellular flame
The growth and wrinkling process of the spherically expanding propane-air flame at
various flame radii is shown in Figure 5. The flame at angle of 270° is first wrinkled
4 W. KIM ET AL.

Figure 2. Original and edge-detected images of large-scale flame.

Figure 3. Flame front evolution at a time interval of 40 ms.

because there is an ignition-spark rod in that direction. Such a wrinkled flame front grows
with time and then a larger wrinkling structure is formed. However, this growth is not a
result of intrinsic instability because the flame is wrinkled by the influence of the rod.
Therefore, in the present study, we focus on the upwardly propagating flame at the range
of 0°–180°. The flame front was nearly flat and stable when flame radius was less than a
critical flame radius associated with the onset of flame acceleration. In our previous study,
the critical flame radius was evaluated by plotting the measured flame speed as a function
COMBUSTION SCIENCE AND TECHNOLOGY 5

0.90

Distance from center [m]


0.85

0.80

0.75

0.70
0 60 120 180 240 300 360
Angle

Figure 4. Detected flame edge in polar coordinates.

360
300
240
Angle

180
120
60
= 62 ms = 122 ms = 182 ms = 242 ms = 302 ms
0
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Flame radius [m]

Figure 5. The growth and wrinkling of spherically expanding flame.

of the flame stretch rate (Kim et al., 2015a). In the case of the propane-air flame of ϕ = 0.8,
the critical flame radius, rc, was found to be rc = 0.24 m, which corresponds to t = 86 ms
(Kim et al., 2015a). The present results exemplify that small-scale cells due to Darrieus–
Landau instability merge and form a bigger cell, as the flame radius increases. The
generated bigger cell grows with time and forms a large single cusp. A single-peak
structure of large amplitude hence emerges. This experimental observation is in qualitative
agreement with previous numerical studies (Gutman and Sivashinsky, 1990; Mukaiyama
et al., 2013).
Figure 6 compares the area, AT, inside the edge of the wrinkled flame, which was
computed by using MATLAB with that of a smooth spherical flame without wrinkling,
defined as AL = π(εSLt)2. Here, ε is the expansion ratio, SL is the laminar burning velocity,
and t is time. SL was calculated by CHEMKIN with the San Diego Mechanism. The value
of SL was 0.3055 m/s. The values of AT and AL were initially close, but after t ≈ 50 ms, AT
deviated from AL because of the formation of a cellular flame. The difference between AT
and AL then increased with time. Shown in Figure 7 is the wrinkling factor, defined as AT/
AL, of the large-scale propane-air flame as a function of Pe-Pec where Pe = r/δ is the Péclet
number with δ being the laminar flame thickness, and Pec = rc/δ is the critical Péclet
number. The wrinkling factor was always greater than 1 (AT/AL > 1), and the values for
Pe-Pec < 0 were greater than those for Pe-Pec > 0 because of the influences of ignition.
Initially, the wrinkling factor decreased for Pe-Pec < 0. This is because the cellular
6 W. KIM ET AL.

2.5

Flame cross-sectional area [m2]


2.0 Exp. (Wrinkled flame)
Cal. (Laminar flame)

1.5

1.0

0.5

0.0
0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35
Time [s]

Figure 6. Comparison between the laminar and wrinkled flame cross-sectional area.

2.8 4

Flame speed [m/s]


2.4
Exp. (Wrinkled flame)
Wrinkle Factor

2
2.0 Cal. (Laminar flame)

1.6

0
-400 0 400 800 1200 1600
c

Figure 7. Evaluated wrinkling factor of large-scale flame and the flame speed as a function of the
Péclet number.

development due to instability at small flame radius was restrained by strong flame stretch,
which is inversely proportional to radius for a smooth spherical flame. As the flame radius
increased, the influence of flame stretch was gradually reduced, and cells on the flame
surface grew by enhanced instability. For Pe-Pec > 0, the wrinkling factor increased with
an increase in Pe-Pec because of the cellular development due to the instability mechan-
ism. These results demonstrate that the cellular structure of large-scale flame is developed
by the reduction of the stretching effect and the enhancement of Darrieus–Landau
instability. This tendency was also observed from comparison between measured flame
speed and calculation defined as Vf = dr/dt = εSL, in Figure 7. The flame speed for Pe-
Pec > 0 increased with an increase in Pe-Pec. It is indicated that the onset of flame
COMBUSTION SCIENCE AND TECHNOLOGY 7

acceleration for Pe-Pec > 0 was appeared due to enhancement of Darrieus–Landau


instability.
The experimental peak-to-peak amplitude is plotted as a function of the average flame
radius in Figure 8. The peak-to-peak amplitude of the wrinkling flame increased with the
flame radius. It is indicated that the flame wrinkling due to Darrieus–Landau instability
grows as time progresses and leads to an increase in the flame area. Based on the results of
linear stability analysis, Bradley (Bradley, 1999; Bradley and Harper, 1994) described the
effects of both Darrieus–Landau and diffusional–thermal instabilities and provided the
amplitude, a, of the perturbation relative to unperturbed flame radius is

a ¼ a0 Rωð1þΩ=Pe ln RÞ (1)

where a0 is the initial dimensionless amplitude, R = r/rc is the scaled flame radius, Pe = r/δ
is the Péclet number where δ is the laminar flame thickness, and ω is a growth rate
parameter that depends on the density ratio ε while the other parameter Ω depends upon
ε, the Lewis number Le, and the Zel’dovich number β. The first term (ω) in the exponent
of the equation expresses the growth rate due to Darrieus–Landau instability, while the
second term (ωΩ/PelnR) is associated with diffusional–thermal instability. The detailed
equations for ω, Ω and β can be found in Refs. (Bradley, 1999; Bradley and Harper, 1994).
This equation illustrates that the amplitude of perturbation depends on Darrieus–Landau
instability as well as diffusional–thermal instability. In the present case, the amplitude of
the perturbation is mainly determined by the growth rate of Darrieus–Landau instability
because the flame is diffusional–thermally stable as Le > 1. The experimental value (ω
+ ωΩ/PelnR) in the present large-scale flame was found to be 0.79, which should reflect
strong effects of first term (ω) and weak effects of the second term. However, this issue is
inconclusive and merits further investigations.

0.1
0.08
Peak-to-peak amplitude [m]

0.06

0.04

Slope = 0.79
0.02

0.5 1 1.5
Flame radius [m]

Figure 8. Peak-to-peak amplitude as a function of flame radius.


8 W. KIM ET AL.

Flame acceleration and fractal dimension


The large single-cusp structure and wrinkling on the front of a sufficiently large flame lead
to the acceleration in the propagation speed. According to previous studies, e.g. (Karlin
and Sivashinsky, 2007; Michelson and Sivashinsky, 1977; Mukaiyama et al., 2013;
Sivashinsky, 1977), the cellular structure of flame front due to Darrieus–Landau instability
has fractal characters. In the present study, the fractal nature of large-scale flame was
investigated by evaluating the fractal dimension. The fractal dimension of self-similar
propagation can be quantified through the power law, namely r = rc + Atα, where A is
experimental constant, α is acceleration exponent correspond to fractal excess d = (α-1)/α
(Gostintsev et al., 1988). The values of α in many investigation was directly evaluated by
the power law and applied many research (Akkerman et al., 2013; Akkerman and Law,
2013; Akkerman et al., 2011; Bychkov and Liberman, 1996; Chaudhuri et al., 2012;
Liberman et al., 2004; Pana and Fursenko, 2008). In addition, the relation of flame
speed, dr/dt = αA1/αr(α−1)/α, instead of the radius was applied to avoid the error due to
the influences of spark ignition (Wu et al., 2013; Kim et al., 2015a).
The equation of propagation speed of a flame wrinkled by cellular instabilities was
derived from the fractal theory (Kim et al., 2015b). The flame speed, Vf, can be
expressed as
 d
Pe
Vf ¼ εSL (2)
Pec

where ε is the expansion ratio, SL is the laminar burning velocity, Pec = rc/δ is the critical
Péclet number for the onset of acceleration, and d is the fractal excess, which is related to
the fractal dimension D3 as d = D3 − 2. This formula indicates that the flame speed
increases in proportion to the Péclet number to the power of d. The critical Péclet number
for the onset of acceleration depends on the characteristic properties of Darrieus–Landau
as well as diffusional–thermal instabilities; an experimentally obtained relationship
between the critical Péclet number and the Markstein number is reported in a previous
study (Kim et al., 2015a).
Figure 9 shows the relation between the dimensionless flame speed (Vf/SL) versus the
Péclet number (dimensionless flame radius). The flame speed increased with the increase
in the Péclet number. The flame fractal excess can be evaluated by the power-law relation-
ship between the flame speed and the Péclet number; the slope in Figure 9 corresponds to
the fractal excess d, and its value was found to be d = 0.24. Hence, the value of flame
fractal dimension D3 for the propane-air mixtures of ϕ = 0.8 is D3 = 2.24, correspond to
α = 1.31. The evaluated dimension D3 = 2.24 is in the range between 2.2 and 2.35, which
covers the fractal dimensions measured in a number of large-scale gas explosion experi-
ments (Kim et al., 2015a; Bradley and Harper, 1994; Wu et al., 2013). However, the
accuracy of evaluating the fractal dimension from a relationship between flame speed and
flame radius is rather limited because the error in measuring instantaneous flame speed is
about ± 15% based on the data shown in Figure 9.
A different method, a box-counting method, is therefore tested to evaluate the fractal
dimension. The fractal dimension D2 of the edge of a flame front such as those shown in
Figure 3 can be estimated as
COMBUSTION SCIENCE AND TECHNOLOGY 9

12

10

Dimensionless flame speed


Slope=0.24

4
100 1000
Peclet number

Figure 9. Dimensionless flame speed versus the Péclet number.

4.0

3.5
= 22 ms
= 102 ms
3.0
= 302 ms
2.5
) R

2.0
Log (

1.5

1.0

0.5

0.0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5
Log ( )

Figure 10. Relation between NR and R at t = 22 ms, 102 ms, 202 ms, and 302 ms.

logðNR Þ
D2 ¼  (3)
logðRÞ

where NR is the number of boxes and R is the size of the boxes. The fractal dimension can
be obtained from the slope of logarithmic plot of NR as a function of R. Figure 10 shows
the relations between log (NR) and log (R) at t = 22 ms, 102 ms, 202 ms, and 302 ms. The
evaluated dimensions were D2 = 1.41, 1.06, 1.19 and 1.18 at t = 22 ms, 102 ms, 202 ms,
and 302 ms, respectively. The fractal dimension D3 of the actual flame surface is related to
that of flame edge D2 as D3 = D2 + 1 if the perturbation on the flame surface is isotropic
(Gostintsev et al., 2004). Figure 11 shows the fractal dimension computed by the box-
counting method as a funtion of the Péclet number.
10 W. KIM ET AL.

2.5
Box-counting dimension
2.4 D = 7/3 [Gostinsev et al., 1988]

2.3

Fractal dimesion
D = 7/3

2.2

2.1

2.0

1.9
0 500 1000 1500 2000 2500
Peclet number

Figure 11. Fractal dimension estimated by the box-counting method as a function of the Péclet
number.

The fractal dimension was initially as large as 2.4 when the flame radius was small, a
value slightly greater than D3 = 7/3 proposed by Gostinsev et al. (1988). This high value
presumably reflects the effect of ignition. After the initial ignition period, the fractal
dimension increased with the increasing Péclet number. However, in this experiment,
the evaluated values of dimension from 2.0 to 2.19 was smaller than the typical value,
D3 = 7/3 (Gostintsev et al., 1988; Sreenivasan and Meneveau, 1986). The increasing
tendency of fractal dimension demonstrates that the range of the Péclet number might
have been in a transition regime toward self-similar propagation and the value dose not
reach that of well-developed self-similar propagation. The existence of the transition
regime to self-similarity was confirmed in previous works (Kim et al., 2015a; Gostintsev
et al., 1988; Kim et al., 2013; Wang et al., 2016). We compared the fractal dimensions D3
obtained by the box-counting method as a function of (Pe–Pec)/Pec with previous evalua-
tions in Figure 12. The dimensions for hydrogen-air and methane-air flames of ϕ = 0.7
were estimated by the relation between the fractal dimension and acceleration exponent α
as D3 = (3α–1)/α. For (Pe–Pec)/Pec < 1.5, the fractal dimension of the propane-air flame of
ϕ = 0.8 was close to that of the methane-air flame of ϕ = 0.7, both of which were smaller
than that of the hydrogen-air flame of ϕ = 0.7 presumably because of the influence of
enhanced diffusional–thermal instability for the hydrogen-air flame. As the flame radius
increased, the fractal dimension also increased and eventually saturated. The transition
regime to self-similarity was observed for (Pe–Pec)/Pec < 6, while the well-developed self-
similar regime was achieved when (Pe–Pec)/Pec > 6–8.
The above comparison indicates that the propane-air flame at ϕ = 0.8 has not yet
reached the well-developed self-similar regime even at the flame radius of r = 1 m. This is
because the propane-air flame is diffusional–thermally stable. The critical Péclet number is
then significantly greater than those under diffusional–thermally unstable conditions,
reducing the value of the scaled Péclet number, (Pe–Pec)/Pec. In order words, observing
COMBUSTION SCIENCE AND TECHNOLOGY 11

2.4

2.3

Fractal dimesion 2.2

H2-air [Kim et al., 2015a]


2.1 CH2-air [Kim et al., 2015a]
C3H8-air

2.0
0 3 6 9 12 15
(Pe - Pec ) / Pec

Figure 12. Fractal dimension as a function of (Pe–Pec)/Pec.

the well-developed self-similarity of a freely expanding spherical flame under such a


condition requires further investigations of sufficiently large-scale flames.

Conclusions
In this study, we investigated the growth and wrinkling of a spherically expanding large-scale
flame. In particular, in order to discuss the structures of flames wrinkled due to Darrieus–
Landau instability, the edge of flame front of a propane-air mixture of ϕ = 0.8, for which Le > 1,
was analyzed. The results demonstrate that the flame front was flat and stable when its radius
was less than a critical value associated with the onset of flame acceleration. As the flame radius
increases, small-scale cells merged and formed a large single cusp. The peak-to-peak amplitude
of flame was evaluated and the amplitude of the wrinkling was found to increase with the flame
radius. The fractal dimension was evaluated by the relation between the speed and the flame
radius. The evaluated dimension D3 = 2.24 was in the typical range between 2.2 and 2.35. The
value, however, contains a relatively large experimental error because of the difficulty in
accurately measuring instantaneous flame speed. The fractal dimension was therefore esti-
mated by another method, the box-counting method. The results showed that the dimension
increased with an increase in the Péclet number. The evaluated values of dimension from 2.0 to
2.19 were smaller than the typical range, suggesting that the flame had not reached the well-
developed self-similar regime even at the flame radius of 1 m. The results demonstrated that
the wrinkled structure of a large-scale flame can be characterized by its fractal dimension and
that the transition into the well-developed self-similar regime exists.

ORCID
Wookyung Kim http://orcid.org/0000-0002-2425-7762
12 W. KIM ET AL.

References
Akkerman, V., Chaudhuri, S., and Law, C.K. 2013. Accelerative propagation and explosion trigger-
ing by expanding turbulent premixed flames. Phys. Rev. E., 87, 023008.
Akkerman, V., and Law, C.K. 2013. Flame dynamics and consideration of deflagration-to-detona-
tion transition in central gravitational field. Proc. Combust. Inst., 34, 1921–1927.
Akkerman, V., Law, C.K., and Bychkov, V. 2011. Self-similar accelerative propagation of expanding
wrinkled flames and explosion triggering. Phys. Rev. E., 83, 026305.
Bradley, D. 1999. Instabilities and flame speeds in large-scale premixed gaseous explosions. Philos.
Trans. Royal Soc. A., 357, 3567–3581.
Bradley, D., and Harper, C.M. 1994. The development of instabilities in laminar explosion flames.
Combust. Flame, 99, 562–572.
Bychkov, V.V., and Liberman, M.A. 1996. Stability and the fractal structure of a spherical flame in a
self-similar regime. Phys. Rev. Lett., 76, 2814.
Chaudhuri, S., Wu, F., Zhu, D., and Law, C.K. 2012. Flame speed and self-similar propagation of
expanding turbulent premixed flames. Phys. Rev. Lett., 108, 044503.
Clanet, C., and Searby, G. 1998. First experimental study of the Darrieus-Landau instability. Phys.
Rev. E., 80, 3867–3870.
Darrieus, G. 1938. Propagation d’un front de flame, unpublished work presented at Paris: La
Technique Moderne and le Congrès de Méchanique Appliquèe.
Dobashi, R., Kawamura, S., Kuwana, K., and Nakayama, Y. 2011. Consequence analysis of blast
wave from accidental gas explosions. Proc. Combust. Inst., 33, 2295–2301.
Gostintsev, Y.A., Fortov, V.E., and Shatskikh, Y.V. 2004. The self-similar law of propagation and fractal
surface structure of the free extending turbulent spherical flame. Dokl. Phys. Chem., 397, 68–71.
Gostintsev, Y.A., Istratov, A.G., and Shulenin, Y.V. 1988. Self-similar propagation of a free turbu-
lent flame in mixed gas mixtures. Combust. Explo. Shock+., 24, 563–569.
Gutman, S., and Sivashinsky, G.I. 1990. The cellular nature of hydrodynamic flame instability.
Physica D: Nonlinear Phenomena, 43, 129–139.
Karlin, V., and Sivashinsky, G. 2007. Asymptotic modelling of self-acceleration of spherical flames.
Proc. Combust. Inst., 31, 1023–1030.
Kim, W., Mogi, T., and Dobashi, R. 2013. Flame acceleration in unconfined hydrogen/air deflagra-
tions by using infrared photography. J. Loss Prevent. Proc., 26, 1501–1505.
Kim, W., Mogi, T., Kuwana, K., and Dobashi, R. 2015a. Self-similar propagation of expanding
spherical flames in large scale gas explosions. Proc. Combust. Inst., 35, 2051–2058.
Kim, W., Mogi, T., Kuwana, K., and Dobashi, R. 2015b. Prediction model for self-similar propaga-
tion and blast wave generation of premixed flames. Int. J. Hydrog. Energy., 40, 11087–11092.
Kim, W., Sato, Y., Johzaki, T., Endo, T., Shimokuri, D., and Miyoshi, A. 2018. Experimental study on self-
acceleration in expanding spherical hydrogen-air flames. Int. J. Hydrog. Energy., 43, 12556–12564.
Landau, L.D. 1944. On the theory of slow combustion. Acta Physicochim, 19, 77–85.
Liberman, M.A., Ivanov, M.F., Peil, O.E., and Valiev, D.M. 2004. Self-acceleration and fractal
structure of outward freely propagating flames. Phys. Fluids., 16, 2476.
Michelson, D.M., and Sivashinsky, G.I. 1977. Nonlinear analysis of hydrodynamic instability in
laminar flames–ll. Numerical experiments. Acta Astronaut, 4, 1207–1221.
Mukaiyama, K., Shibayama, S., and Kuwana, K. 2013. Fractal structures of hydrodynamically
unstable and diffusive-thermally unstable flames. Combust. Flame., 160, 2471–2475.
Okafor, E.C., Nagano, Y., and Kitagawa, T. 2016. Experimental and theoretical analysis of
cellular instability in lean H2-CH4-air flames at elevated pressures. Int. J. Hydrog. Energy.,
41, 6581–6592.
Pana, K.L., and Fursenko, R. 2008. Characteristics of cylindrical flame acceleration in outward
expansion. Phys. Fluids., 20, 094107.
Sivashinsky, G.I. 1977. Nonlinear analysis of hydrodynamic instability in laminar flames–I.
Derivation of basic equations. Acta Astronaut., 4, 1177–1206.
Sreenivasan, K., and Meneveau, C. 1986. The fractal facets of turbulence. J. Fluid. Mech., 173, 357–386.
COMBUSTION SCIENCE AND TECHNOLOGY 13

Wang, J., Xie, Y., Cai, X., Nie, Y., Peng, C., and Huang, Z. 2016. Effect of H2O addition on the flame
front evolution of syngas spherical propagation flames. Combust. Sci. Technol., 188, 1054–1072.
Williams, F.A. 1985. Combust. Theory, Westview Press, U.K.
Wu, F., Jomaas, G., and Law, C.K. 2013. An experimental investigation on self-acceleration of
cellular spherical flames. Proc. Combust. Inst., 34, 937–945.

You might also like