You are on page 1of 18

Combustion and Flame 162 (2015) 4084–4101

Contents lists available at ScienceDirect

Combustion and Flame


journal homepage: www.elsevier.com/locate/combustflame

Formation and evolution of distorted tulip flames


Huahua Xiao∗, Ryan W. Houim, Elaine S. Oran
Department of Aerospace Engineering, University of Maryland, College Park, MD 20742, USA

a r t i c l e i n f o a b s t r a c t

Article history: The development and evolution of tulip and distorted tulip flames in closed channels were simulated by
Received 5 May 2015 solving the fully compressible reactive Navier–Stokes equations using a high-order numerical method and a
Revised 25 August 2015
single-step Arrhenius model for the reactions and energy release in a stoichiometric mixture of hydrogen and
Accepted 27 August 2015
air. Important features of the simulations include (1) the development and propagation of acoustic waves and
Available online 18 September 2015
their effects on flame evolution, (2) the formation and collapse of flame cusps, both at the flame front and
Keywords: near the sidewalls, and the effects of cusp collapse on flame propagation, and (3) the appearance of adverse
Premixed flame pressure gradients at the onset of the tulip or a distorted tulip flame, which result in reverse flow in the
Distorted tulip flame unburned gas. The simulations highlight the coupling between pressure waves, adverse pressure gradients,
Flame-flow interaction boundary layers, and the propagating flame front. Whereas the formation of the tulip flame can be attributed
Flame-pressure wave interaction to several effects (such as pressure waves, vortex motion and Landau–Darrieus instabilities), the onset of the
Rayleigh–Taylor instability
distorted tulip flame is strongly influenced by the Rayleigh–Taylor instability.
Boundary layer
© 2015 The Combustion Institute. Published by Elsevier Inc. All rights reserved.

1. Introduction propagating tulip flame. (See, for example, the first figure in [16]).
Tulip flames were first observed photographically [17], where it was
The understanding of the dynamics of premixed flames propagat- shown that these structures develop in closed tubes when the as-
ing in tubes is important in a wide range of combustion processes and pect ratio of the tube is greater than two. The transition to a tulip
applications, including gas explosions in confined regions and energy flame is also accompanied by a sudden decrease in flame propaga-
production by internal combustion engines [1–12]. A premixed flame tion speed. After a tulip flame forms, it generally propagates down
propagating in a tube is actually a complicated, dynamic process, in- the tube until combustion is complete. In very long tubes, for exam-
volving all of the complexities of ignition, development of a laminar ple, tubes with aspect ratios greater than 20, tulip flames collapse at
flame, and the subsequent interactions of this flame with boundary some point, again become convex toward the unburned mixture, and
layers and pressure wave under changing background conditions. The then finally develop again into a tulip flame [18]. In addition, more
initial laminar flame that can evolve from a small spark is intrinsically recent experimental work has shown the development of multiple
unstable due to a variety of hydrodynamic and combustion instabili- tulip flame shapes at a flame front [19,20].
ties, such as Darrieus–Landau (DL), thermal-diffusive, and Rayleigh– There have been many investigations of the physical process
Taylor (RT) Instabilities [2,13,14]. The flame may accelerate quickly leading to formation of a tulip flame. To date, however, there is no
and undergo a series of changes in its shape. It may then further decisive, single explanation of the physical mechanism by which they
evolve into a turbulent flame, and deflagration-to-detonation tran- form. Various elements of existing explanations include interactions
sition (DDT) is even possible, depending on the reactive material and of the flames with pressure waves [18], effects of viscosity and flame
confinement geometry [1,2,10,15]. quenching [4,17], hydrodynamic instabilities [16,21–23], vortex
In this paper, we examine the development of an initially laminar structures forming in the burned gas [3,24–26], and Rayleigh–Taylor
flame that becomes a “tulip flame,” a term that qualitatively describes instabilities [5].
the flame shape, and subsequently becomes a “distorted tulip flame” Now we focus on the behavior of a premixed hydrogen–air flame,
(or DTF). The initial laminar flame may be convex, due to the con- initially at atmospheric pressure and temperature, propagating in a
finement of the tube and interaction with boundary layers ahead of closed tube. Information from fully compressible two-dimensional
the flame. Subsequently, this shape may invert to become a concave (2D) numerical simulations is combined with results of prior ex-
periments [20] to give a detailed description of the evolution of the
flame-front structure. We describe the formation of the flame into

Corresponding author. a tulip flame, and the subsequent development of distorted and
E-mail address: hhxiao@umd.edu, huahuahsiao@gmail.com (H. Xiao). multiple tulip flames. All of these complex flame structures evolve

http://dx.doi.org/10.1016/j.combustflame.2015.08.020
0010-2180/© 2015 The Combustion Institute. Published by Elsevier Inc. All rights reserved.
H. Xiao et al. / Combustion and Flame 162 (2015) 4084–4101 4085

from a single small ignition. It is hoped that descriptions of the duced in the equivalence ratio range 0.84 ≤ φ ≤ 4.22 in premixed
physical processes, derived from the computations, will help us to hydrogen–air (corresponding to hydrogen concentration 26–64% by
understand the interactions among the flame front, fluid dynamics, volume in air). The distortions, which appeared as secondary cusps,
boundary layers, and pressure waves. originated in the vicinity of the tips of the primary tulip flame lips
and moved toward the primary cusp as the flame evolved. (This is
2. Background discussed and shown more clearly in text below). The initiation of a
DTF is consistent with a sudden decrease in both the speed of the
Here we review a selected group of the many papers on tulip leading flame front and the growth rate of the pressure. The authors
flames, focusing on the work we believe is most relevant to the DTF. suggested that these phenomena, which were accompanied by flame
In the process, we note those results that seem to be contradictory oscillations, may be due to the strong coupling between the flame
and that show that we do not completely understand the controlling and pressure waves. Another speculation is that cellularity due to
mechanisms. thermal-diffusive instabilities might be of minor importance for the
Clanet and Searby [5] conducted an experimental and analytical formation and evolution of DTFs, since the Lewis number for a mix-
study of tulip-flame evolution. From their analysis, they concluded ture that generated a DTF ranged from approximately 0.9–2.2 [37].
that neither acoustic effects nor boundary layers are dominant in the The mechanism of the DTF formation remains unresolved. Previ-
tulip flame formation. They also suggested that the RT instability was ous numerical studies [19,38] could not resolve enough of the possi-
important, but this mechanism has not been fully supported by their ble controlling phenomena because they used numerical algorithms
or any later theory. Subsequent numerical simulations, all using zero- that were both too diffusive with inadequate resolution. As a re-
Mach-number models [16,24,27], which could not take acoustics into sult, they could not provide enough detail to analyze the interac-
account, showed that tulip flames can be produced without including tions among the flame front, pressure waves and boundary layers. Al-
the effects of pressure waves. In addition, these studies concluded though a second DTF was observed in the previous experiments [38],
that tulip flame formation is sensitive to system parameters, such as until now we did not have simulations that would allow us to explore
the geometry of the combustion chamber, boundary conditions at the how the flame evolves after the second DTF, and then through to the
walls, and the mixture composition. final stages of combustion.
Gonzalez et al. [28] carried out a parametric study of tulip flame What is needed for further discussion of DTFs is an understand-
formation in an enclosure by means of extensive 2D numerical sim- ing of the interaction between the flame fronts, boundary layers, and
ulations with a “thickened-flame” model. They concluded that the pressure waves generated by the flame as it propagates down the
transverse velocity gradient and DL instability are two crucial mecha- tube, and how these interactions might affect flame instabilities. As
nisms for the development of tulip flames. As these authors indicated, noted above [20], it was suggested that pressure waves might play
their simulations suffered from severe inaccuracies in representing an important role in the formation of the DTF, possibly because of
the viscosity terms. They also suggested that wall friction is not im- a RT instability. There is, however, neither a rigorous demonstration
portant for the formation of a tulip flame. In juxtaposition to this, nor a detailed examination of the interaction between the DTF and
Marra and Continillo [27] argued that wall friction is the determin- pressure waves. For example, it is not clear how pressure waves are
ing cause for the initiation of a tulip flame. Their conclusions were generated, propagate, and play a role in flame instabilities.
based on computations that solved the full Navier–Stokes equations, The purpose of this paper is to study the dynamics and evolution
again, assuming zero-Mach number. It is interesting that these mod- of DTFs using unsteady, fully compressible solutions of the Navier–
els, which did not include acoustic effects, were able to produce tulip Stokes equations with accuracy and resolution sufficient to resolve
flames. pressure waves and their interactions with flames and boundary lay-
It has also been shown that the vorticity field is important. In the ers. These results are then described and compared to prior experi-
process of tulip flame formation, vortices created near the flame front ments.
in the burned gas were reported in many experimental and numerical
studies [6,19,29–31]. Matalon and Metzener [3,26] proposed a math-
ematical theory to support the explanation that the tulip flame is 3. Physical and numerical models
caused by vortical motion in the burnt gas. Nevertheless, an inviscid
numerical simulation with restrictions to irrotational flow described Now we describe the physical and numerical models, and then
by Dunn-Rankin et al. [16] showed the formation of a tulip flame in compare the numerical results to the theoretical predictions [39]
the absence of vorticity effects. Vortices can also be generated by a and the prior experimental observations [20]. The current numeri-
curved flame [32]. For example, a pair of vortices in the burnt region cal model has been extensively used in a variety of combustion and
can be created by the cusp of a curved flame arising from DL instabil- explosion problems [10,40–46]. Comparisons shown below provide
ity. Although the onset of flame-front inversion that leads to a tulip further tests and demonstration of its reliability and accuracy for pre-
flame may be related to the DL instability [21,33], the tulip flame can- dicting transient combustion processes.
not be explained simply by linear stability analysis of DL instability The 2D computational domain considered in Fig. 1 shows a d ×
[3,26]. l cm closed rectangular channel. Note that only half of the channel
Bychkov et al. [34], following the work of Clanet and Searby [5], is simulated. Extensive prior tests comparing full and half channel
suggested an analytical model for the acceleration of a finger-shaped
flame and the initiation of a tulip flame in the early stages of laminar
flame propagation in open long tubes. They found that the accelera-
tion of the finger-shaped flame and tulip flame formation do not de-
pend on Reynolds number. For fast flames propagating in these tubes,
such as those of hydrogen–oxygen flames, compressibility can play a
role in the flame acceleration and tulip flame evolution [35].
Recently, an additional complication was added by the discovery
of the distorted tulip flame [19,20,36]. These structures appear as ad-
ditional cusps on the lips of the original tulip flame. Under some con-
ditions, a second DTF formed on the lips before the first DTF collapsed Fig. 1. Computational domain. All walls are adiabatic with no-slip reflecting
[20]. The experiments [19,20,36] demonstrated that DTFs can be pro- boundaries.
4086 H. Xiao et al. / Combustion and Flame 162 (2015) 4084–4101

situations showed that the results for this laminar flame problem
were extremely symmetrical.

3.1. Numerical model

The numerical simulations solve the 2D fully compressible


Navier–Stokes equations with a model of chemically reacting stoi-
chiometric mixture for hydrogen and air. The governing equations
that describe the reacting flow are written as [10,47]
∂ρ ∂(ρ u) ∂(ρv)
+ + = 0, (1)
∂t ∂x ∂y
∂(ρ u) ∂(ρ u2 + p) ∂(ρ uv) ∂σxx ∂σyx
+ + = + , (2)
∂t ∂x ∂y ∂x ∂y
∂(ρv) ∂(ρv2 + p) ∂(ρ uv) ∂σyy ∂σxy
+ + = + , (3)
∂t ∂y ∂x ∂y ∂x

∂(ρ E ) ∂ [u(ρ E + p)] ∂ [v(ρ E + p)]


+ +
∂t ∂x ∂y
∂(uσxx + vσxy − qx ) ∂(vσyy + uσyx − qy )
= + + qω̇, (4) Fig. 2. Variation of pressure with time at the center of the right end wall of a 1 cm ×
∂x ∂y 7 cm tube. The minimum cell size is dxmin .
   
∂ρY ∂(ρ uY ) ∂(ρvY ) ∂ ∂Y ∂ ∂Y
+ + = ρD + ρD −ω̇,
∂t ∂x ∂y ∂x ∂x ∂y ∂y where A and Ea are the pre-exponential factor and activation energy,
respectively. This type of simplified chemistry model has been exten-
(5)
sively tested and used to solve various combustion problems includ-
ing laminar flame propagation [40,41], cellular detonations [42,43],
4 ∂ u 2 ∂v 4 ∂v 2 ∂u flame-shock interactions and DDT [10,44–46].
σxx = μ − μ , σ = μ − μ , (6)
3 ∂ x 3 ∂ y yy 3 ∂ y 3 ∂ x The governing equations are solved using using a fifth-order
MUSCL algorithm [48] with HLLC fluxes [49,50]. The unsteady for-
∂u ∂v
σxy = σyx = μ +μ , (7) mulations are solved using a three-stage, third-order explicit Runge–
∂y ∂x Kutta scheme. Adaptive mesh refinement (AMR) using the Paramesh
∂T ∂T library [51] provides local mesh refinement in the region of impor-
qx = −K , q = −K , (8) tant features of the flow and combustion, such as flame fronts, strong
∂x y ∂y
pressure waves, and boundary layers. The flame is ignited by a small
ρ RT semi-circular pocket of hot, burned gas with a radius of 0.5 mm at
p= , (9) the left end wall. The center point of the ignition zone is on the tube
M
axis. No-slip adiabatic boundary conditions are used at the walls of
p 1 2 the tube, so that we can see the effects of boundary layers and their
E= + (u + v2 ), (10)
(γ −1)ρ 2 growth.
where ρ , t, u, v, p, T, E, q, ω̇, R, M, and Y are the density, time, Numerical resolution tests were performed by varying the value
stream-wise velocity, transverse velocity, pressure, temperature, of dxmin , the minimum cell size used in the most refined layer of
total energy, heat release, reaction rate, universal gas constant, mix- the computational grid. Figure 2 shows the variation of pressure with
ture molecular weight, and mass fraction of the unburned mixture, time recorded at the center of the right end wall using three different
respectively. The quantities σ xx , σ yy and σ xy are the components minimum grid sizes, dxmin = 1/160, 1/320 and 1/640 cm, for a 1 cm ×
of deviatoric stress tensors. And qx and qy are the heat fluxes. The 7 cm tube. It is found that dxmin = 1/640 cm (15.625 μm), correspond-
diffusive fluxes in Eq. (5) are described by Fick’s law of diffusion. The ing to 22 computational cells in the flame itself at initial conditions, is
specific heat capacity at constant pressure cp and the ratio of specific adequate for the present calculations. The tests also show that dxmax
heat capacities γ are taken as constant. Note that the total energy = 1/40 cm is sufficient.
in Eq. (10) contains no chemical energy and the specific heats are
constant, so that the contribution of species diffusion to the heat 3.2. Modeling parameters
flux vector is not included in Eq. (8). The transport properties, i.e.,
the dynamic viscosity μ, mass diffusivity D, and thermal diffusivity The parameters chosen to simulate a stoichiometric premixed
k = K/(qc p ), are calculated as a function of temperature, hydrogen–air flame are given in Table 1 [44]. These parameters were
taken from Gamezo et al. who did extensive simulations and com-
μ = μ0 T n , D = D0 T n , k = k0 T n , (11)
parisons to experiments performed a posteriori. The calculated prop-
where μ0 , D0 , k0 are the transport properties under initial condi- erties of a one-dimensional (1D) steady-state laminar flame for the
tions. The temperature index is assumed to be a constant, n = 0.7. system are also shown in the table. The flame thickness is defined
We assume that the mixture is an ideal gas, so that the specific heat here as the distance between mass fraction iso-surfaces Y = 0.1 and
capacity at constant pressure is given by c p = γ R/(γ −1). Y = 0.9. Note that the Lewis number is Le = 1 and the flame is thus
The reaction rate is taken into account using a single-step diffusively stable.
Arrhenius-type chemistry kinetics, The chemistry model can reproduce the major properties of a
 −E  hydrogen–air mixture and is computationally affordable for cal-
ω̇ = dY/dt = AρYexp a
, (12) culating the multidimensional transient combustion in the system
RT
H. Xiao et al. / Combustion and Flame 162 (2015) 4084–4101 4087

Table 1 Table 2
Input model parameters and calculated properties for simulating a stoichiometric Comparison between Zelcovich’s theory and the computation.
hydrogen–air flame.
pmax (kPa) Tumax (K) Tbmax (K) Tbmin (K)
Input parameters
Theoretical 847.89 405.76 2961.0 2282.39
p0 1.0 atm Initial pressure Numerical 842.59 406.72 2822.97 2240.02
T0 298.0 K Initial temperature
ρ0 0.86 kg/m3 Initial density
γ 1.17 Specific heat ratio
M 21.0 kg/kmol Molecular weight process is complete due to non-equilibrium isentropic compression
A 6.85 × 109 m3 /kg-s Pre-exponential factor of the burnt gas. The maximum and minimum temperatures occur at
Ea 46.37RT0 Activation energy the very end of combustion
q 43.28RT0 /M Chemical heat release  
μ0 = D0 = k0 2.9 × 10−6 kg/s-m-K0.7 Initial transport parameters ε−1
Tbmax = T0 ε γ −1/γ 1+ , (15)
Output flame properties γ
Su0 3.11 m/s Laminar burning velocity  
Tb 2140.65 K Post-flame temperature ε−1
ρb Tbmin = T0 ε γ −1/γ + . (16)
0.12 kg/m3 Post-flame density γ
δL 0.34 mm Laminar flame thickness
The parameters for comparison between the theoretical prediction
and the computational results are summarized in Table 2. Overall, the
considered here. These properties include the laminar flame speed, computed parameters agree reasonably well with those predicted by
laminar flame thickness, adiabatic flame temperature, transport the Zeldovich’s theory.
coefficients, sound speed, Chapman–Jouget (CJ) detonation velocity, There are, however, differences to be expected between a theory
detonation cell size and half reaction-zone thickness. The model based on solely an energy conservation principle and a fully com-
was calibrated [10] by (1) performing a series of 1D calculations of pressible solution of the Navier–Stokes equations. The difference be-
flame and detonation structures for a range of input parameters, tween the computation and theory for Tbmax is relatively large (about
(2) choosing a set of parameters that most closely reproduces the 138.0 K). There are at least two reasons. The first is that thermal dif-
properties of both a laminar flame and a detonation, (3) conducting fusion broadens the peak of the temperature profile in the computa-
2D computations using the model parameters chosen in step (2) for tions. The second is that the acoustic waves, in this numerical solu-
calculating detonation cell size, (4) performing full simulations of tion, contain energy.
flame propagation and DDT using these parameters, and (5) com- In the theory, the pressure is assumed to be uniformly distributed
paring these results to experiments. The 2D and 3D hydrogen–air in the vessel. Numerical simulations show that the pressure oscillates
flame propagation and DDT in obstructed channels using this model during the combustion process, as shown in Fig. 2. The pressure pmax
[44,45] reproduces the main DDT phenomena observed in exper- in Table 2 is the time-averaged value of the pressure recorded after
iments. That is, it predicts flame acceleration, various combustion the combustion is complete. The amplitude of the pressure oscillation
instabilities, choking flames, quasi-detonations, detonations, and at the end of the burning process is about 40.0 kPa, and the tempera-
stochastic behavior of detonation initiation. ture oscillates accordingly. Following the isentropic relationship, the
An recent investigation by Taylor et al. [52] shows that the flame- resulting temperature oscillation amplitude should be less then 0.7%.
shock interaction and detonation initiation in hydrogen–air mixtures Thus the deviations of Tumax , Tbmax and Tbmin arising from the pressure
calculated using this simplified model are in general agreement with oscillation are within 2.75 K, 19.11 K and 15.16 K, respectively.
those using a detailed chemical reaction mechanism. In this study, we
use this model for calculating hydrogen–air flame propagation in the 3.4. Comparison to prior experiments
enclosure described by Fig. 1.
In the prior experiments [19,20], the combustion chamber was a
3.3. Comparison to Zeldovich’s theory 53 cm-long closed rectangular tube with a cross-section of 8.2 cm ×
8.2 cm. The tube was filled with premixed stoichiometric hydrogen in
Zeldovich [39] proposed an analytical theory to predict explosion air and ignited with a single spark gap near the left wall on the tube
parameters near the end of gas explosion in an closed vessel. These axis. The evolution of the flame front was recorded with high-speed
parameters include the maximum pressure in the entire burning pro- schlieren cinematography.
cess, and the maximum temperature of the unburned gas when com- Figure 3 a shows the development of the tulip flame and the DTF
bustion is extremely close to completion, and the maximum and min- in these experiments. Figure 3b shows density gradients taken from
imum temperatures of the burnt gas when combustion is completed. numerical simulations. Both the experiment and the simulation show
In the theory, the maximum pressure in an enclosure achieved that the flame shape first changes from convex to concave, and then
when combustion is complete is to a tulip flame, and then DTFs. The DTF develops into a triple tulip
flame as the secondary cusps approach the center of the primary tulip
pmax = (γ − 1)qρ0 + p0 . (13)
flame lips (at 7.8 ms in Fig. 3a and 9.033 ms in Fig. 3b). A second
The basic assumptions leading to this equation is that initial mixture DTF is generated with a cascade of distortions superimposed on the
is an ideal gas with a constant specific heat ratio, and it undergoes primary lips (at 9.0 ms in Fig. 3a and 10.571 ms in Fig. 3b) before the
adiabatic compression and energy is conserved. The maximum tem- disappearance of the first DTF. Thus the key features of the flame-
perature of the unburned mixture Tumax occurs close to the end of front evolution in the experiment are reproduced in the numerical
the combustion process and is computed based on the isentropic re- simulation. In particular, the shapes of the calculated DTFs reproduce
lationship as those in the experimental observations.

Tumax = T0 ε γ −1/γ , (14)


4. Results
where ε = pmax /p0 . In addition, the region where the mixture is
burnt first has the highest temperature and the region where the ma- The numerical simulations discussed in detail here were per-
terial is burned last has the lowest temperature after the combustion formed for a 4 cm × 28 cm tube, approximately half the size of the
4088 H. Xiao et al. / Combustion and Flame 162 (2015) 4084–4101

Fig. 3. Schlieren images of the premixed hydrogen–air flame front in (a) the experiment [20] and (b) the simulation. Note that the entire experimental domain in the vertical
direction is not shown since both the upper and lower edges of the schlieren window are 0.133 cm away from the lower and upper tube walls. The numerical schlieren images are
shown for the same spatial proportion of the computational domain. The actual heights of the domains shown in the images in the experiment and numerical simulation are 7.93
and 7.74 cm, respectively. The numerical images were mirrored on the symmetry line.

experiment discussed above. Nevertheless, the flame evolution in this these stages, we comment briefly on the effects of ignition and the
case is qualitatively similar to the experimental results. overall behavior of the pressure waves generated by the ignition, as
shown in the computational schlieren.
4.1. Overall evolution of the flame The ignition process creates pressure waves that propagate out
ahead of the flame. The first pressure waves are reflected from the
Figure 4 shows the computational schlieren and corresponding sidewalls, and these reflections propagate down the tube. As seen
temperature maps at selected times. These show that the develop- from the schlieren image, the reflections form a series of criss-crossed
ment of the tulip flame and the DTF can be discussed in terms of five lines of more intense local pressure. They are not, however, strong
stages: enough to be called a shock train, though they have that form. The
first pressure waves bounce back once they arrive at the right wall,
(1) a spherical flame,
and the reflection first reaches the flame at about 1.4 ms. Once the
(2) a finger-shaped flame,
reflections reach the flame front, they are reflected back into the un-
(3) a flame with its skirt touching sidewalls
burned gas as well as transmitted through the flame into the burned
(4) a tulip flame, and
region. There is, however, no noticeable effect on the flame shape due
(5) distorted tulip flames.
to these weak acoustic interactions.
These stages were also discussed earlier in relation to the exper- After ignition, the flame front develops into a hemispherical shape
iments [19]. The “flame skirt” refers to the parts of the flame front which expands for a short time, and, by 0.49 ms begins to evolve into
which are moving toward the sidewalls. Before discussing each of a finger shaped flame. In this first stage, the flame expands outwards
H. Xiao et al. / Combustion and Flame 162 (2015) 4084–4101 4089

Fig. 4. Sequences of schlieren and temperature fields showing the numerical premixed hydrogen–air flame propagating in a 4 cm × 28 cm closed tube.

freely, without effects of the sidewalls. Because of the no-slip bound- flame front. Once the expansion wave reaches the leading tip of the
ary conditions, boundary layers develop along the walls as the flame flame, the flame suddenly decelerates, and it continues to decelerate
propagates, as also shown in an earlier study [40]. throughout the third stage as the sides of the flame are interacting
In stage (2), the finger-shaped flame stage, the flame is signifi- with the walls. During this stage, expansion waves are continuously
cantly elongated along the tube axis. The flame elongation is primar- generated from the sidewalls due to the reduction of flame surface
ily caused by gas expansion under confinement of the tube sidewalls, area.
as demonstrated by Kurdyumov and Matalon [53] who studied flame The flame starts to invert at around 2.72 ms, after which a tulip
acceleration in long, narrow, open channels. Meanwhile, the flame flame is formed (see 3.05 and 3.31 ms in Fig. 4). The flame starts to
close to the left wall is inclined and makes an acute angle with the become distorted at about 3.91 ms after a pronounced tulip flame
wall due to boundary layer effects, as shown 1.37 ms. As the bound- has been established. Secondary cusps are produced near the original
ary layer forms, the flow velocity changes, increasing from zero at the tulip flame lips (4.14 ms), which was also observed in the experiments
wall. The nonuniform flow stretches the flame, and different portions [19,20]. The secondary cusps move toward the center of the tube and
of the flame move into different background flow environments, so form a triple tulip flame, as shown at 4.38 ms. The first DTF even-
that the flame shape resembles the flow velocity profile. tually disappears, and this is accompanied by the collapsing of the
The flame first touches the sidewalls near the left corners of the primary tulip cusp. The collapsing cusp generates a relatively strong
tube at about 1.69 ms (not shown). This interaction generates pres- pressure wave, which can be seen in the collapse region at 5.81 ms.
sure waves due to the sudden loss of flame area, which had been pro- This collapse process will be clarified below.
viding a source of energy driving flame expansion. As we will discuss A second DTF forms before the full collapse of the first one, as
and show in more detail below, this interaction of the flame with the shown, for example, at 5.44 ms. These cusps behave in a similar way
wall results in a weak rarefaction wave. as those of the first DTF. Before the disappearance of the second DTF,
The flame skirt forms small angles with the sidewalls, and the however, a third DTF forms, as seen at 7.03 ms. In fact, there are a
flame evolves to the point where the skirt is almost parallel to the series of DTFs, one forming after the other until the leading tip of
sidewalls. At 1.95 ms (not shown), the flame skirt contacts the side- the flame comes very close to the end of the tube. Near the end of
walls for a second time, and the flame surface area is reduced more the flame propagation, the flame develops into a classical tulip shape
sharply than it was the first time. The result is a much stronger ex- again, with small wrinkles spreading over the entire flame front, as
pansion wave, as shown by the schlieren image at 1.96 ms. This ex- shown at 9.71 ms. The flame maintains this shape until combustion
pansion wave propagates at the sound speed and soon overtakes the is completed.
4090 H. Xiao et al. / Combustion and Flame 162 (2015) 4084–4101

Fig. 5. Schlieren images, pressure fields, and the generation of expansion waves after the flame contacts the sidewalls.

The schlieren images also show noticeable vortex motion in the point means that there is a sudden loss of expansion of combustion
burnt gas after the tulip flame forms. As the flame continues to product. This subsequently generates the semi-circular rarefaction
evolve, vortical structures develop, one after another behind the waves shown at the bottom and top left corners at 1.70 ms in Fig. 5.
flame front. These vortices travel in the opposite direction of the These pressure waves are reflected as they reach the sidewalls and
flame front, forming a street of vortices in the burnt region. The lead- propagate toward the flame front. The pressure waves catch up with
ing vortex takes the appearance of a mushroom structure, as shown the leading tip at about 1.76 ms and cause the flame to decelerate. The
in the schlieren images (e.g., at 7.03 and 8.63 ms). As the flame lead- disappearance of the flame in the left corners, which occurs because
ing front approaches the right end wall of the tube, the leading vortex the flame has burned out, also generates weak expansion waves.
also moves closer to the left end wall due to compression of the com- The second contact of the flame skirt with the sidewalls results in
bustion products, as shown at 9.71 ms. a further reduction of the flame surface area and generates stronger
expansion waves due to the very shallow angles between the flame
4.2. Collapsing flame cusps and generation of pressure waves skirt and the sidewalls. This is as shown at 1.95 and 1.96 ms in Fig. 5.
These pressure waves overtake the leading tip at about 2.02 ms and
One interesting feature that occurs during the DTF propagation cause dramatic flame deceleration. This is discussed further in below
is related to the behavior of flame cusps. In this section, we describe in terms of variations in flame-velocity profiles.
the evolution and characteristics of flame cusps to provide a more de- Figures 6 and 7 show selected schlieren and pressure fields during
tailed exploration of the flame dynamics. This description also helps the formation of the tulip flame and the first DTF, respectively. As the
us to understand how pressure waves are generated, propagate, and tulip flame grows, slits (or flat gaps) containing unburned gas form
interact with the flame front. between the flame front and the tube sidewalls, as shown at 3.10 ms
Figure 5 shows computational schlieren images and correspond- in Fig. 6. The shape of these gaps means that the flame later contacts
ing pressure fields and waves generated as the flame develops into the sidewalls over an area that is actually in front of the existing con-
the third flame stage, that is, when the flame skirt touches the tube tact point. In other words, the contact point appears to move jump
sidewalls. Note that adiabatic boundary conditions are applied on the ahead discontinuously! This process reduces the flame surface area in
tube walls, so that the contact of the flame with the sidewalls math- a very short time, and so generates the noticeable expansion waves
ematically resembles the collapsing of a cusp. After the flame has shown at 3.31 ms in Fig. 6. The same phenomenon is also observed
reached the sidewalls, the reduced flame surface area at the contact during the development of the DTF, as show in Fig. 7. Nevertheless,
H. Xiao et al. / Combustion and Flame 162 (2015) 4084–4101 4091

Fig. 6. Schlieren images, pressure fields, and generation of rarefaction waves during the tulip flame evolution.

Fig. 7. Schlieren, corresponding pressure fields and pressure wave generation during the development of the first distorted tulip flame.

the pressure waves generated are not as strong as those generated in caused by these expansion waves is not as significant. The phenom-
the tulip flame growth process. ena of cusp collapsing and pressure wave generation occurs continu-
Figure 8 shows schlieren and pressure fields during the evolution ally until the end of the combustion.
of the second and third DTFs. As the first DTF disappears, a small re-
gion of unburned gas is created in the cusp, as can be seen at 5.44 ms 4.3. Flame dynamics
in Fig. 4. When the cusp collapses, a circular expansion wave is gener-
ated, as shown at 5.8 ms in Fig. 8. Simultaneously, a tiny pocket of un- In this section, we discuss the flame dynamics in terms of the
burned gas is left behind. (This can be seen a s small black dot behind leading tip of the flame, the flame front along the centerline, and
the cusp along the centerline at 5.80 ms.) Another expansion waves the changes in pressure. The flame front along the centerline is the
is generated when this small burning pocket collapses, as shown at same as the flame leading tip before flame inversion. After inver-
5.81 ms. During the formation of the third DTF, flat gaps of unburned sion, the leading edge of the flame near the sidewalls moves ahead
gas form near the sidewalls, as shown at 6.09 ms. The collapsing of of the flame at the centerline. Figure 9 shows the evolution of the po-
these near-wall flame structures also generates the weaker expan- sition and the propagation speed of the leading tip of the flame. Here
sion waves shown at 6.17 ms. As shown below, flame deceleration the position is defined as the distance from the flame location to the
4092 H. Xiao et al. / Combustion and Flame 162 (2015) 4084–4101

Fig. 8. Schlieren, corresponding pressure fields, and generation of expansion waves during the evolution of the second and third distorted tulip flames.

position at the end wall where ignition originally occurred. The flame flame front starts to invert, as shown in Fig. 4. The flame surface area,
propagation speed is the displacement speed of flame front in the however, stops decreasing at a later time, about 2.92 ms, so there is
laboratory frame of reference. The flame-tip propagation speed is also a time lag of about 0.2 ms. This time lag is caused by the interaction
related with the variation of flame surface area, as shown in Fig. 10. between the flame front and pressure waves reflected from the right
The flame propagation speed fluctuates greatly throughout the end wall.
duration of the combustion process, similar to the results in the prior In fact, at the beginning of the third stage, the reduction in the
experiments [19,20]. The flame accelerates quickly after ignition due flame surface area precedes the decrease of the flame speed, since
to the fast increase of the flame surface area accompanying the sec- the expansion waves need time to overtake the flame front. These ex-
ond finger-like stage. The flame decelerates first at about 1.4 ms when pansion waves are reflected and become compression waves as they
it interacts with pressure waves that were generated since ignition reach the right end wall. The flame decelerates further when it en-
and then reflected from the right end wall. At 1.69 ms, pressure waves counters these waves.
were generated by the contact of the flame skirt with the side walls. Both the flame surface area and flame propagation speed fluctuate
These pressure waves reach the flame front at 1.76 ms and this re- together during the rest of the flame propagation. Although the flame
sults in a second flame deceleration. This deceleration is small since speed is closely related to the flame surface area, their phase differ-
the pressure waves causing it are weak. The flame decelerates for the ence changes with time due to the interaction between the flame and
third time at about 2.02 ms, and this is caused by expansion waves pressure waves as well as the increase of the sound speed. In particu-
generated by the second contact of the flame skirt with the sidewalls lar, the flame surface area increases rapidly before the end of combus-
at about 1.95 ms. This flame deceleration is larger and persists for a tion. The reason for this is that the flame again develops into a normal
longer time. The deceleration terminates at around 2.72 ms when the tulip flame with a slender cusp as it approaches the right end wall, as
H. Xiao et al. / Combustion and Flame 162 (2015) 4084–4101 4093

Fig. 11. Flame position and propagation (displacement) speed along the centerline as
Fig. 9. Position and propagation (displacement) speed of the leading tip of the flame
a function of time.
as a function of time.

Fig. 12. Pressure rise and pressure growth rate inside the tube as a function of time.

Fig. 10. Flame surface area and the propagation (displacement) speed of the leading
tip of the flame as a function of time.
with pressure waves. The amplitude of the pressure oscillation in-
creases very quickly when the flame approaches the right end indi-
shown at 9.71 ms in Fig. 4. Moreover, the flame front is considerably cating that the pressure waves are significantly amplified at the later
wrinkled due to the decrease of the flame thickness and the growth stages of burning.
of intrinsic instabilities as the pressure inside the tube increases.
The location and propagation speed of the flame front along the 5. Discussion
centerline of the channel as a function of time is shown in Fig. 11. In
particular, note the large jumps in the flame speed along the center- The dynamics of a premixed flame in an enclosed tube is influ-
line. These are the result of collapsing cusps. Towards the end of the enced by the background flow induced by combustion, which in-
combustion process, the cusp jumps more frequently, since deeper cludes the formation and destruction of boundary layers. Pressure
cusps are produced and the flame front becomes more corrugated (as waves and flame instabilities can be also of major importance for
shown in Fig. 4). The maximum propagation speed in the x−direction the development of the tulip and distorted tulip flames. Here we dis-
resulting from the collapse can be as high as 1500 m/s, primarily due cuss the interactions between the flame front and these physical phe-
to the cusp collapse. Another reason for this is the extremely shallow nomena and the subsequent effects on the flame deformations. As
angle between the flame front of cusp and the centerline, since the noted earlier, the processes that create both tulip and distorted tulip
flame orientation is almost tangential to the axis (e.g., the flame at flames are not conclusively determined. In previous studies of pre-
5.8 ms in Fig. 8). This surprising result explains the experimental ob- mixed flame propagation in a closed duct [19,38], the characteristic
servation of the steep increase of the primary tulip cusp speed in the length scales of the flame (such as the flame thickness) and bound-
previous work [19]. ary layer (such as boundary-layer thickness) were not resolved, espe-
Figure 12 shows the pressure and its growth rate as a function of cially in the later stages of the calculations. The result was that the
time as recorded on the centerline at the right end wall. These quanti- flame speed and pressure rise were underpredicted, at least in the
ties oscillate also, which further reflects the strong coupling of flame late stages of the computations. In the present work, the flame front,
4094 H. Xiao et al. / Combustion and Flame 162 (2015) 4084–4101

Fig. 13. Time sequence of instantaneous streamlines and the pressure field near the flame front during the development of the tulip flame. The heavy white line indicates the
location of flame front (iso-surface Y = 0.5). The unit of pressure is kPa.

boundary layer and pressure waves are resolved, so that we now have concave flame front that appears after the deceleration of a finger-
a more accurate picture of the interaction between the flame and shaped flame.
these phenomena. The interactions between a flame front and its surroundings, in-
cluding the near-field flow and pressure gradients, have direct effects
5.1. Formation of the tulip flame on the flame dynamics. Figure 13 shows a sequence of numerical im-
ages of the pressure field and flow patterns near the flame front when
As mentioned in references [19,34,35], the concept of tulip flame the tulip flame forms. In the flame acceleration stages, the flow both
can be ambiguous, since the word “tulip” is used to designate any in the burnt and unburned regions mainly moves in the same direc-
concave flame with a cusp pointing towards the burnt gas. This sit- tion as the flame front. This reflects the expansion of the hot-gas re-
uation can arise from a number of very different physical phenom- gion and the induced motion ahead of the flame.
ena, e.g., the interaction of flame with shock wave [54], the DL in- After the flame skirt touches the side walls, vortices are generated
stability [55], and the inversion subsequent to sudden deceleration in the burnt region near the sidewalls, e.g., at 2.26, 2.37, 2.59 and
of finger-shaped flame [3–6,19]. The tulip flame discussed here is the 2.66 ms in Fig. 13. These vortices expand in time as they approach
H. Xiao et al. / Combustion and Flame 162 (2015) 4084–4101 4095

and move through the flame front. During this process, the flame the flow at the flame lips is positive (in the right direction) while
front starts to invert, as shown at 2.73 ms. The reverse flow (in the flow at the cusp remains reversed. This reinforces the cusp, as shown
left direction) occupies the near-field burnt and unburned regions, at 3.91 ms in Fig. 4. The far-field burnt region maintains vortical
as shown at 2.81 and 2.88 ms. In particular, an important feature of motions during the entire process of tulip-flame evolution.
this flow is that the reversed flow comes not only from the burned Matalon and Metzener [3,26] proposed a mathematical theory
region, but also from the far-field unburned gas, as shown at 2.73 ms. to support the explanation that the tulip flame is caused by vortex
This observation is similar to the result by Kurdyumov and Matalon motion behind the flame. According to their theory, the curved flame
[53], who found that adverse pressure gradient drives unburned shape generates strong vorticity which then accumulates due to
gas towards opposite direction. The reverse flow in the unburned the confinement. A pair of vortices is created by the highly curved
gas passes through the flame front, joining together with the burnt segments of the flame near the sidewalls before the flame front
reverse flow, and forming a strong reverse flow at center around the flattens. The intense vortices eventually advect the flame into a tulip
flame front, as shown at 2.81 ms. This flow pattern facilitates devel- shape. Numerical simulations by Dunn-Rankin et al. [16], based on
oping tulip flame. After the tulip flame is initiated (2.95–3.39 ms), a inviscid, zero-Mach number model, suggests that a tulip-shaped

Fig. 14. Flow pattern and pressure field around the flame front during the evolution of the first distorted tulip flame. The heavy white line denotes the location of flame front
(iso-surface Y = 0.5). The unit of pressure is kPa.
4096 H. Xiao et al. / Combustion and Flame 162 (2015) 4084–4101

Fig. 15. Streamline and pressure field around the flame front during the evolution of the second distorted tulip flame. The heavy white line designates the location of flame front
(iso-surface Y = 0.5). The unit of pressure is kPa.

flame could be produced by the DL instability alone without the are connected with the pressure wave and boundary layer, as will be
presence of vorticity. Nevertheless, the time for their tulip flame to discussed below. A similar flow pattern in the vicinity of flame front
form is considerably longer than that in their experiment. Numerical may be seen in [53].
study by Gonzalez et al. [28] also indicates that the time scale of tulip
formation due to the DL instability is much larger than that due to 5.2. Formation of the distorted tulip flames
the confined flow field generated by combustion itself in a closed
tube. Although the DL instability could produce a tulip-shaped flame, The interactions between the flame front, reverse flow and vortex
the results of our simulations support the mechanism of vortex motion both at the tube center and near the side walls create con-
motion described by Matalon and Metzener. ditions favorable for the initiation and formation of the tulip flame.
Vortices are also generated ahead of the flame front near the side- Here we show that they are also favorable for the initiation of the
walls, as shown at 2.73 ms in Fig. 13. In addition, flat gaps of unburned distorted tulip flames.
material form between the tulip flame lips and the sidewalls due to Figures 14 and 15 show near-field streamlines and pressure fields
the effects of these vortices, as seen at 2.95 and 3.17 ms near the walls during the evolution of the first and second DTFs. By 3.76 ms in Fig. 14,
in Fig. 13 (also see Fig. 6). As the tulip flame develops further, these a well-pronounced tulip flame has formed and there does not appear
structures disappear. The flow pattern and generation of the vortices to be any vortical flow near the flame front. The flow in the unburned
H. Xiao et al. / Combustion and Flame 162 (2015) 4084–4101 4097

gas is positive (in the right direction), as shown at 3.63 and 3.76 ms. The flows reverses in the unburned gas just ahead of the locations
Just before the formation of the first DTF (e.g., at 3.88 ms), vortices of the secondary cusps, as shown at 4.01 ms. This flow in the sec-
are generated close to the tips of the tulip flame in unburned gas near ondary cusps stays reversed until a well-pronounced DTF is formed,
the sidewalls. As the height of the vortices increases, the first DTF be- as shown at 4.14, 4.26 and 4.39 ms. This flow decelerates the sec-
gins to form, as shown at 4.01 ms. The motion of these vortices move ondary cusps and is strong enough to make the cusps move back-
the flame front from near the sidewalls towards the centerline. These wards, as shown at 4.26 ms. The combined effects of the vortex mo-
vortices create flat gaps between the flame front and sidewalls in the tion ahead of the flame front near the sidewalls, the positive flow in
manner that we saw for the tulip flame. This effect may also acceler- the vicinity of the primary cusp, and the reverse flow near the sec-
ate the flame tip near the sidewalls and consequently contribute to ondary cusps promote the onset and development of the first DTF.
the start and growth of the distorted tulip shape. The vortices decay The flow pattern during the development of the second and the
and ultimately disappear as a well-pronounced DTF forms, as shown succeeding DTFs, shown in Fig. 15, is very similar to that of the first
at 4.26, 4.39 and 4.51 ms. DTF. That is, vortices are generated next to the flame tips just before
At 4.01 ms, the far-field flow ahead of the flame is reversed. In the initiation of the second DTF (4.87 and 4.93 ms). These vortices
contrast to what was seen as the original tulip flame formed (Fig. 13), tend to push the flame tips toward the tube center, again creating flat
this reverse flow does not penetrate the flame front. Instead, the near- gaps of unburned gas between the flame and sidewalls. Vortical mo-
field flow becomes positive, as seen, for example, at 4.14, 4.26 and tion in the burnt gas is generated immediately behind the primary
4.39 ms. Meanwhile, there are two vortices produced closely behind cusp and behaves in a similar manner to that in the formation of the
the primary tulip cusp shortly after the onset of the DTF, and these first DTF. A reverse flow is also observed at the initiation locations of
grow quickly with the development of the DTF, as shown at 4.01, 4.14 the second DTF. Finally, as the second DTF forms, reverse flow reoc-
and 4.51 ms. These vortices remain immediately behind the primary curs in the unburned region. There are, however, also some notable
cusp and cause positive flow in the burnt region in the tube center, as differences in the flow patterns between the first and second DTFs. In
shown 4.26, 4.39 and 4.51 ms. The unburned gas flow confined in the particular, note the differences in when and where in the DTF forma-
primary tulip cusp remains positive while the DTF is forming. These tion process the flow is reversed and the shapes and strengths of the
vortices help to accelerate the primary cusp forward. vortices formed.

Fig. 16. Vertical pressure profiles at (a) x = 1.6 cm and (b) 11.3 cm at different times immediately after the second and third sudden contact of flame front with the tube sidewalls,
respectively.

Fig. 17. Horizontal pressure distributions at different times at (a) y = 0 cm and (b) −1.6 cm during the initiation of the tulip flame and the first distorted tulip flame, respectively.
4098 H. Xiao et al. / Combustion and Flame 162 (2015) 4084–4101

A detailed discussion of the differences between the distorted 5.4. The role of pressure waves
tulip flame and the classical tulip flame and other relevant flame
phenomena was given in [19]. The distorted tulip flame also differs The pressure waves in the system were generated by several dif-
in principle from the multiple formation of classical tulip flame re- ferent processes, e.g., ignition, contact of flame front with the tube
ported by Guenoche [18]. In his experiments, after the formation of walls (mimicking cusp collapse), and the collapsing flame cusps off
a classical tulip flame, the flame inversion can reverse itself and the the walls. These pressure waves, which were shown to be expansion
flame front again becomes convex towards the unburned gas. In suffi- waves, are reflected by the tube walls and flame front and propagate
ciently long tubes, e.g., a tube with an aspect ratio of 24, this inversion back and forth across the tube. Generally, pressure waves interact
process can repeat itself several times until the end of combustion. with a flame and cause instabilities, as discussed in the experimen-
tal [58,59], numerical [60–62] and analytical [63] studies. Here we
5.3. Cusp collapse look more carefully at the role of the pressure waves in the hope that
this will give us insight into their interactions with the combustion-
The phenomenon of a collapsing cusp accompanied by a sub- generated flow, boundary layers and flame front.
sequent increase in the local flame speed was also found in nu- Figure 16 shows the vertical pressure distribution at different
merical simulations of premixed turbulent flames [56,57]. In both times at x = 1.6 cm and 11.3 cm at selected times. These locations
situations, the structure of the reaction zone of at the flame surface are the points at which expansion waves were created by the second
remain essentially that of a planar laminar flame. In the case of the (see Fig. 5) and third (see Fig. 6) contacts of the flame front with the
turbulent flame, the turbulent cascade failed to penetrate the inter- tube sidewalls. The pressure waves generated at these two locations
nal flame structure, and thus the action of small-scale turbulence was are relatively strong, as shown in Section 5.2. As discussed above, a
suppressed throughout most of the flame. In our case, where there is small pressure spike was created at the very beginning of the flame
no turbulent spectra, we might expect some changes in the laminar contact with the sidewall, as shown at 1.946 ms (a) and 3.301 ms (b)
flame as the local pressure increases, but the structure remains intact. in Fig. 16. This implies that a small explosion occurred as the flame
Note the small explosions at the very beginning of cusp collapse touched the sidewall. That is, the contact of a flame with an adiabatic
shown in Fig. 5. This leads to a pressure spike that occurs before the wall mimics the collision of two flames. Nevertheless, the pressure
expansion wave is formed. The pressure spike arises due to the fo- decreases rapidly with the sudden loss of flame surface area after the
cusing of a thermal flux from hot products into colder fuel. This local flame has touched the sidewalls. This process is accompanied by the
increase in temperature heats the reactants and increases the rate at generation of expansion waves.
which they burn, thereby causing the two flame sheets to propagate The pressure waves generated in this way propagate spherically
faster toward each other. This effect of a cusp collapsing and causing outward and gradually decrease in amplitude. The initial amplitude
a sudden increase in flame speed was also reported by Poludnenko of the pressure waves generated at 1.6 cm and 11.3 cm are 2.65 kPa
et al. [56,57] when the cusps collapsed in the turbulent flame brush. and 2.09 kPa, respectively. According to Teerling et al. [60] and

Fig. 18. Baroclinic torque field at different times during the initiation of the tulip flame. The unit of baroclinic torque is 1/s2 .
H. Xiao et al. / Combustion and Flame 162 (2015) 4084–4101 4099

Fig. 19. Baroclinic torque field at different times in the formation process of the first distorted tulip flame. The unit is 1/s2 .

Shalaby et al. [61], a pressure wave of amplitude on the order of 1 kPa and reverse flow began in the near-field, unburned gas, as shown at
or higher can lead to strong flame wrinkling and periodic changes 2.66 and 2.73 ms. As the pressure peak passed through the flame front
of the flame shape through a RT instability. The initial amplitude from the unburned to the burnt side, the reverse flow penetrated the
of a pressure wave in our numerical simulation is of this critical flame front, as shown at 2.81 and 2.88 ms in Fig. 13. Vortices ahead
size. Although the amplitude of the wave decreases with time, the of the flame were countered and dissipated by strong backward flow.
combined effects of many pressure waves on the curved flame could The tulip flame began during this process.
be significant, given that these pressure waves may coalesce to form After the tulip flame started to form, a pressure wave reflected
pressure waves with sufficiently large amplitude during the transient from the ignition end wall, and its peak passed through the flame
burning process. front from the burnt to unburned side. This positive pressure gradient
Figure 17 shows the horizontal pressure and the corresponding pushed the flow in the unburned gas toward the right end wall, creat-
profiles of the mass fraction of unburned gas at different times at ing a positive flow in the unburned gas, as shown at 2.95 and 3.17 ms.
y = 0 cm and −1.6 cm during the initiation of the tulip flame and the The same type of phenomena regarding the interactions between
first DTF, respectively. The horizontal line y = 0 cm corresponds to the pressure waves, flow, and the flame front were seen in the formation
vertical location where the tulip flame originates, and y = −1.6 cm is of the first and the subsequent DTFs, as shown in Figs. 14 and 15.
close to the vertical location where the first DTF forms. As shown in During the initiation of the tulip flame and DTFs, vortices were
Fig. 17, a pressure wave moves from left to right side of the domain formed near the sidewalls ahead of the flame front by the adverse
with an adverse pressure gradient before the onset of both the tulip pressure gradient in the boundary layer. The adverse pressure gra-
flame and the first DTF. After the flame begins to deform, the pressure dient first induced reverse flow in the boundary layer. Then the
gradient changes sign. The difference between the high and low pres- reverse flow interacted with the adjacent positive flow, as shown, for
sure can be as large as 17 kPa for the tulip flame and 18 kPa for the example, at 2.59 ms in Fig. 13. This interaction caused rapid vorticity
first DTF. This is large enough to initiate RT instabilities on a curved production. This mechanism of interaction between the pressure
flame during the initiation process of the tulip flame and the first DTF. waves and the boundary layer is consistent with that reported by Ott
Pressure fields near the flame front for the tulip flame, and the first et al. [40].
and succeeding DTFs were shown above in Figs. 13, 14 and 15, respec- A curved premixed flame interacting with an acoustic wave is sus-
tively. There we noted that the flame decelerate abruptly just before ceptible to RT instability. The RT instability is usually driven by forces
the tulip flame or a DTF was initiated. Before the tulip flame formed, associated with the differing gas densities on the two sides of the
a pressure wave traveled from the right side toward the convex flame flame front. There is a strong misalignment of the density gradient
front with an adverse pressure gradient (in an opposite direction to and pressure gradient in a curved flame when exposed to a pressure
the flame propagation), as shown at 2.37–2.66 ms in Fig. 13. When the wave. This misalignment results in a baroclinic torque, as shown in
pressure peak was close to the flame front, vortices started to appear Figs. 18 and 19, which generates vorticity. As a result, strong vortical
4100 H. Xiao et al. / Combustion and Flame 162 (2015) 4084–4101

structures are generated and these modify the flow near the flame Therefore, the pressure waves have a role in the development of the
front. tulip flame, although they are not the only contributing factor.
The interactions of the flame and the pressure waves are taking Vortices near the sidewalls and ahead of the flame front were also
place in a flow field that is already quite complicated by many prior observed in the initiation process of DTFs when the strong adverse
interactions, as shown in Figs. 13, 14 and 15. Even during the tulip pressure gradient passes though the flame front. The vortex motion
flame formation process, there can be effects of baroclinic torque due advects the flame front near the sidewalls towards the tube center.
to the presence of pressure gradients, as shown Figs. 13 and 18. There- Moreover, additional vortices form immediately behind the primary
fore, even if pressure waves are not the determining factor for tulip tulip cusp and result in positive flow in the burnt gas along the cen-
flame formation, the effect of the flame-pressure wave interaction terline which moves the primary cusp forward. The pressure waves
can at least promote tulip flame development in the closed tube con- reverse the flow just before the formation of DTF. In the meantime, a
sidered here. reverse flow is generated ahead of the locations where the secondary
For the DTFs, there is no obvious vortical motion just behind the cusps form, thus promoting the growth of these cusps. Finally, the
primary tulip lips, as shown in Figs. 14 and 15. In this case, the ef- DTF is formed by the combined effects of the near-field flow motions,
fect of RT instabilities arising from the interaction between the flame i.e., the reverse flow near the initiation locations of the secondary
front and acoustic waves is important for the formation of the DTFs. cusps, the vortices ahead of the flame front near the sidewalls, and
In addition, the vortex motions behind the primary cusp and ahead of the vortical motion closely behind the primary tulip cusp.
the flame front next to the sidewalls conspire with the RT instabilities Our conclusion is that pressure waves have a significant influence
to cause the development of the DTFs. on the flow evolution and thus play an essential role in the DTF for-
mation. Their interaction with the curved flame front has a destabi-
6. Summary and conclusions lizing effect and facilitate the formation of a DTF by triggering unsta-
ble modes, i.e., Rayleigh–Taylor instabilities. The formation of a DTF
This paper presented numerical simulations of the formation and therefore can be thought of as being the result of Rayleigh–Taylor in-
propagation of tulip and distorted tulip flames in closed tubes. The stability driven by pressure waves.
two-dimensional reactive Navier–Stokes equations coupled with the
ideal gas equation of state and a single-step Arrhenius chemistry Acknowledgment
model were solved using a fifth-order MUSCL algorithm and adap-
tive gridding to capture important features of the flow and reaction This work was supported in part by the University of Mary-
waves. land through Minta Martin Endowment Funds in the Department
In general, the numerical simulations showed flame dynamics of Aerospace Engineering, and through the Glenn L. Martin Institute
similar to those observed in prior experiments and provided details Chaired Professorship at the A. James Clark School of Engineering.
of the process through which the flame front evolves from a convex The authors would like to thank Professor Jinhua Sun for his valu-
flame, to a tulip flame, to a distorted tulip flame (DTF), and then de- able discussions and support. The authors also thank the support pro-
velops succeeding distortions. A series of DTFs forms, one after an- vided by National Natural Science Foundation of China (Projects Nos.
other, before the flame reaches the end of the tube. The simulations 51406191 and 51376174) and the Chinese Postdoctoral International
also showed that the initiation of tulip flame and DTFs occurred just Exchange Program (2013). Parts of this work were funded by the Of-
after the flame underwent a sudden deceleration. The formation of fice of Naval Research (project No. N00014-14-1-0177). The authors
these flames, and the increased flame surface area associated with acknowledge the University of Maryland supercomputing resources
them, resulted in subsequent flame acceleration. (http://www.it.umd.edu/hpcc) made available in conducting the re-
One important result observed in the computations is the col- search reported in this paper.
lapse of flame cusps and the “half flame cusps” that appear as flat
References
gaps between the flame front and the sidewalls. The collapse gener-
ates pressure waves, which were shown to be expansion waves. The [1] E.S. Oran, V.N. Gamezo, Combust. Flame 148 (2007) 4–47.
collapse also results in an abrupt increase in the flame-propagation [2] S.B. Ciccarelli, G. Dorofeev, Prog. Energy Combust. Sci. 34 (2008) 499–550.
speed along the tube centerline. These pressure waves, together with [3] M. Matalon, P. Metzener, J. Fluid Mech. 336 (1997) 331–350.
[4] R. Starke, P. Roth, Combust. Flame 66 (1986) 249–259.
the pressure waves generated by the early contacts of the skirt of the
[5] C. Clanet, G. Searby, Combust. Flame 105 (1996) 225–238.
finger flame with the sidewalls, interact with the flame and cause the [6] D. Dunn-Rankin, R.F. Sawyer, Exp. Fluids 24 (1998) 130–140.
flame speed and pressure growth rate to oscillate. The process of cusp [7] M. Fairweather, G.K. Hargrave, S.S. Ibrahim, D.G. Walker, Combust. Flame 116
formation and collapse, accompanied by flame accelerations and de- (1999) 504–518.
[8] G.M. Abu-Orf, R.S. Cant, Combust. Flame 122 (2000) 233–252.
celerations, repeats until the combustion is finished. [9] D. Bradley, M. Lawes, K. Liu, Combust. Flame 154 (2008) 96–108.
Large-scale vortices appear just behind the flame skirt near the [10] D.A. Kessler, V.N. Gamezo, E.S. Oran, Combust. Flame 157 (2010) 2063–2077.
tube sidewalls after the flame skirt touches the sidewalls. The vortices [11] S.B. Dorofeev, Proc. Combust. Inst. 33 (2011) 2161–2175.
[12] V.N. Kurdyumov, M. Matalon, Proc. Combust. Inst. 34 (2013) 865–872.
expand with time and overtake the flame front. Reverse flow forms [13] T. Poinsot, D. Veynante, Theoretical and numerical combustion, 2, Edwards RT
and dominates the near-field region due to the vortex motion. This Inc., Philadelphia, 2005.
process creates conditions that allow the flame to invert. In addition, [14] M. Matalon, Proc. Combust. Inst. 32 (2009) 57–82.
[15] M. Kuznetsov, V. Alekseev, I. Matsukov, S.B. Dorofeev, Shock Waves 14 (2005)
during the initiation of the tulip flame, an adverse pressure gradi- 205–215.
ent is imposed on the flame front. The pressure gradient reverses the [16] D. Dunn-Rankin, P.K. Barr, R.F. Sawyer, Proc. Combust. Inst. 21 (1986) 1291–1301.
flow in the unburned region and creates vortices ahead of the flame [17] O.C.d. C. Ellis, Fuel Sci. 7 (1928) 502–508.
[18] H. Guenoche, G.H. Markstein, Nonsteady Flame Propagation, Pergamon Press,
front near the sidewalls by interacting with the boundary layers. The New York, 1964, p. 107.
adverse pressure gradient and the large-scale vortical motion behind [19] H. Xiao, D. Makarov, J. Sun, V. Molkov, Combust. Flame 159 (2012) 1523–1538.
the flame front conspire to generate a strong reverse flow at center [20] H. Xiao, Q. Wang, X. Shen, S. Guo, J. Sun, Combust. Flame 160 (2013) 1725–1728.
[21] M. Matalon, J.L. Mcgreevy, Proc. Combust. Inst. (1994) 1407–1413.
around the flame front, and consequently cause the formation of tulip
[22] B. N’konga, G. Fernandez, H. Guillard, B. Larrouturou, Combust. Sci. Technol. 87
flame. In addition, the vortices ahead of the flame front facilitate the (1992) 69–89.
initiation of the tulip flame. These vortices are also responsible for [23] J.W. Dold, G. Joulin, Combust. Flame 100 (1995) 450–456.
the flat gaps between the flame front and sidewalls. The amplitude [24] D.A. Rotman, A.K. Oppenheim, Proc. Combust. Inst. (1986) 1303–1310.
[25] A.K. Kaltayev, U.R. Riedel, J. Warnatz, Combust. Sci. Technol. 158 (2000) 53–69.
of the pressure waves generated by contact of the flame front with [26] P. Metzener, M. Matalon, Combust. Theor. Model. 5 (2001) 463–483.
the sidewalls is high enough to cause Rayleigh–Taylor instabilities. [27] F.S. Marra, G. Continillo, Proc. Combust. Inst. 26 (1996) 907–913.
H. Xiao et al. / Combustion and Flame 162 (2015) 4084–4101 4101

[28] M. Gonzalez, R. Borghi, A. Saouab, Combust. Flame 88 (1992) 201–220. [46] T. Ogawa, V.N. Gamezo, E.S. Oran, J. Loss Prev. Process Ind. 26 (2013) 355–362.
[29] D. Dunn-Rankin, R.E. Sawyer, Proceedings of the 10th ICDERS, 1985, pp. 115–130. [47] R.W. Houim, K.K. Kuo, J. Comput. Phys. 230 (2011) 8527–8553.
Berkley, California. [48] B. Thornber, A. Mosedale, D. Drikakis, D. Youngs, R.J.R. Williams, J. Comput. Phys.
[30] T. Kratzel, E. Pantow, M. Fischer, Int. J. Hydrogen Energy 23 (1998) 45–51. 227 (2008) 4873–4894.
[31] K. Kuzuu, K. Ishii, K. Kuwahara, Fluid Dyn. Res. 18 (1996) 165–182. [49] A. Harten, P.D. Lax, B. van Leer, SIAM Review 25 (1983) 25–61.
[32] V. Bychkov, M.A. Liberman, Phys. Rep. 325 (2000) 116–237. [50] E.F. Toro, M. Spruce, W. Speares, Shock Waves 4 (1994) 25–34.
[33] J. Chomiak, G. Zhou, Proc. Combust. Inst. 26 (1996) 883–889. [51] P. MacNeice, K.M. Olson, C. Mobarry, R.d. Fainchtein, C. Packer, Comput. Phys.
[34] V. Bychkov, V. Akkerman, G. Fru, A. Petchenko, L.E. Eriksson, Combust. Flame 150 Commun. 126 (2000) 330–354.
(2007) 263–276. [52] B.D. Taylor, R.W. Houim, D.A. Kessler, V.N. Gamezo, E.S. Oran, AIAA Paper (2013)
[35] D.M. Valiev, V. Akkerman, M. Kuznetsov, C.K. Eriksson L. E. Law, V. Bychkov, Com- 1171.
bust. Flame 160 (2013) 97–111. [53] V.N. Kurdyumov, M. Matalon, Proc. Combust. Inst. 35 (2015) 921–928.
[36] H. Xiao, Q. Wang, X. He, J. Sun, X. Shen, Int. J. Hydrogen Energy 36 (2011) 6325– [54] G.H. Markstein, Proc. Combust. Inst. 6 (1957) 387–398.
6336. [55] V.V. Bychkov, S.M. Golberg, M.A. Liberman, L.E. Eriksson, Phys. Rev. E 54 (1996)
[37] Z. Chen, M.P. Burke, Y. Ju, Proc. Combust. Inst. 32 (2009) 1253–1260. 3713–3724.
[38] H. Xiao, W. An, Q. Duan, J. Sun, Int. J. Hydrogen Energy 38 (2013) 12856–12864. [56] A.Y. Poludnenko, E.S. Oran, Combust. Flame 157 (2010) 995–1011.
[39] Y.B. Zeldovich, G.I. Barenblatt, V.B. Librovich, G.M. Makhviladze, The Mathemati- [57] A.Y. Poludnenko, E.S. Oran, Combust. Flame 158 (2011) 301–326.
cal Theory of Combustion and Explosions, 1, Consultants Bureau, New York, 1985. [58] G. Searby, D. Rochwerger, J. Fluid Mech. 231 (1991) 529–543.
[40] J.D. Ott, E.S. Oran, J.D. Anderson, AIAA J. 41 (2003) 1391–1396. [59] G. Searby, Combust. Sci. Technol. 81 (1992) 221–231.
[41] V.N. Gamezo, E.S. Oran, AIAA J. 44 (2006) 329–336. [60] O.J. Teerling, A.C. McIntosh, J. Brindley, V.H.Y. Tam, Proc. Combust. Inst. 30 (2005)
[42] V.N. Gamezo, A.A. Vasil’ev, A.M. Khokhlov, E.S. Oran, Proc. Combust. Inst. 28 1733–1740.
(2000) 611–617. [61] H. Shalaby, K.H. Luo, D. Thvenin, Combust. Flame 161 (2014) 2868–2877.
[43] A.M. Khokhlov, J.M. Austin, F. Pintgen, J.E. Shepherd, AIAA Paper (2004) 0792. [62] M. Gonzalez, Combust. Flame 107 (1996) 245–259.
[44] V.N. Gamezo, T. Ogawa, E.S. Oran, Proc. Combust. Inst. 31 (2007) 2463–2471. [63] V. Bychkov, Phys. Fluids 11 (1999) 3168–3173.
[45] V.N. Gamezo, T. Ogawa, E.S. Oran, Combust. Flame 155 (2008) 302–315.

You might also like