Você está na página 1de 13

Effect of Wall Heat Losses and Blockage Ratio on Flame Propagation in

Obstructed Micro-Chambers
Orlando Ugarte, Sinan Demir, Berk, Demirgok, Vyacheslav Akkerman*
Center for Alternative Fuels, Engines and Emissions (CAFEE)
Computational Fluid Dynamics and Applied Multi-Physics Center (CFD&AMP)
Department of Mechanical and Aerospace Engineering, West Virginia University
Vitaly Byhkov
Department of Physics, Umea University, Sweden
Damir Valiev
Department of Applied Physics and Electronics, Umea University, Sweden
I. Introduction
Deflagration-to-detonation transition (DDT) is one of the most important and fascinating
combustion phenomena with wide range of applications from pulse-detonation engines to safety
issues such as the prevention of mining accidents [1-3]. The reason and the key element of DDT
is spontaneous flame acceleration in tubes and channels from a low laminar flame speed to
nearly sonic values, which implies ultra-sonic propagation in the reference frame of the tubes
walls by the end of the process [4-6]. Two major mechanisms of flame acceleration in channels
have been elucidated: the Shelkin mechanism in the case of smooth walls [7,8] and the
mechanism of ultra-fast flame acceleration in channels with obstacles [9]; the present work
focuses on a geometry with smooth walls.
The most typical geometry of DDT in experiments and energy-production devices
corresponds to a relatively long channel, with the flame propagating from the closed channel end
to the open one. According to the Shelkin mechanism, a flame front in channels with smooth
walls accelerates because of the thermal expansion of the burning gas and the non-slip boundary
conditions at the walls. Expansion of the burning gas produces a flow of the fuel mixture, which
becomes non-uniform due to the non-slip boundary conditions. The non-uniform flow makes the
flame front curved, which increases the burning rate and creates a positive feedback between the
flame and the flow, hence leading to the flame acceleration. Although the qualitative idea of
flame acceleration was suggested by Shelkin already in the 1940s [7,8], the quantitative theory
of the process has been developed and supported by extensive numerical simulations only
recently, by Bychkov et al. [10,11]. Among other results, the theory [10,11] predicted powerful
exponential flame acceleration in micro-channels (at least at the initial, almost isobaric stage of
the process), and a decrease of the scaled acceleration rate with the channel width characterized
Re S L R /
SL
R
by the Reynolds number
, where
is the laminar flame speed,
is the channel

half-width (radius), and


is the kinematic viscosity. The theoretical predictions [10,11] have
been supported by later experiments on DDT in micro-channels with diameters about 1 mm and
below [12,13]. At the same time, the experiments [12,13] have demonstrated some features of
flame dynamics different from the theoretical predictions [10,11], such as a linear regime of
flame acceleration instead of the exponential one. Reference [5] has explained the difference by
the influence of gas compression effects ignored in [10,11], which moderate the acceleration
process and make it linear for sufficiently large values of the flow Mach number, see also Ref.
[14] for the theoretical explanation of this effect. By the end of the acceleration process, the

flame propagation speed saturates to the Chapman-Jouguet deflagration speed [15-17], unless an
explosion of the fuel mixture happens earlier.
Thermal losses to the walls, which are inevitably present in any experiment on DDT, is
another effect expected to work against the flame acceleration mechanism. For example, in
channels with isothermal cold walls, the thermal losses eventually cool the burnt gas down to the
wall temperature, thus reducing effectively the effect of thermal expansion as the key element of
the flame acceleration mechanism. In spite of the obvious importance of the thermal losses, quite
surprisingly, not a single theoretical work on DDT has yet considered the influence of this effect
on flame acceleration. At the same time, in the presence of strong thermal losses, one may
question the very possibility of flame acceleration and DDT in channels with cold walls. Indeed,
one should naturally expect a dominating role for thermal losses in sufficiently narrow channels,
which may not only stop flame acceleration, but even lead to complete extinction of the burning
process. In wide channels the role of thermal losses decreases, but the flame acceleration
mechanism becomes weaker too with the increase of the Reynolds number [10,11], and it
remained unclear which of these two effects prevails.
In the present work we investigate the influence of thermal losses on flame acceleration in
channels with cold isothermal walls. The problem is solved by using direct numerical
simulations of the complete set of the Navier-Stokes combustion equations. The investigation
can be split into the analyses of the flame dynamics in two configurations. First, we consider
tubes and channels including thermal boundary conditions by keeping the wall at a constant
temperature. Second, the same configuration with the inclusion of obstacles at the chamber sides
is studied.
II. Basic equations and description of the computational approach
The numerical simulations are performed by an in-house program that solves the set of mass,
momentum and energy conservation equations, accounting for transport processes. The set of
chemical reactions is approximated by a single irreversible one-step reaction, obeying the
Arrhenius law. Moreover, the burnt and unburnt gases are set to obey the ideal gas behavior.
The code is based on a cell-centered, finite volume scheme. It is second order accurate in
time, fourth order in space for the convective terms, and second order in space for the diffusive
terms. The solver is adapted for parallel computations and is available in Cartesian (both twoand three-dimensional) and cylindrical axisymmetric versions, with a self-adaptive structured
grid. It has been widely used in combustion research (e.g. [9-14]), including validation with
exact solutions and experimental data.
The set of mass, momentum, energy and species equations solved by the program reads:
1


r ur
uz 0
t r r
z

u r r u r2 r ,r u r u z r , z P 0
t
z
r
r r

(1)

u z 1 r u r u z r , z u z2 z , z P 0
t
z
z
r r

r P u r r ,r u r r , z u z q r P u z z , z u z r , z u r q z
t r r
z

(2)

(3)

, (4)

Y 1 r ui Y Y u z Y Y Y exp E a R pT
t
Sc r z
Sc z
R
r r

0
where

and 1 for 2D and axisymmetric geometries respectively,

QY CV T u z2 u r2
2

(5)

(6)
Q

Y
is the total energy per unit volume,
the mass fraction of the fuel,
the energy release from
CV
qi
the reaction, and
the heat capacity at constant volume. The energy diffusion vector
is
C T Q Y
C T Q Y

q r P

q r P

Pr r Sc r
Pr z Sc z
,
.
(7)
i, j
The stress tensor
takes the form
u i u j 2 u k

i, j

i, j
x j xi 3 x k

(8)
0
1
in the 2D configuration (
), while in the axisymmetric geometry (
) it reads
u r u z
2 u r u z u r
2 u z u r u r
2
2

r ,r


z,z


r , z

3 r
z
r
3 z
r
r
r
z
,
,
. (9)
Finally, the last term in Eq. (2) takes the form
2 u r u r u z
2

3 r
r
z
(10)
0
1
0

Sc
Pr
if
, and
if
. Here
is the dynamic viscosity,
and
the Prandtl and
Schmidt numbers, respectively. The fuel-air mixture and burnt gas are assumed to be perfect
CV 5R p / 2m C P 7 R p / 2m
m 2.9 10 2 kg/mol
gases of constant molar mass
, with
,
, and
P R p T / m
RP 8.31 J mol K
the equation of state
, where
is the universal gas constant.
We consider a one-step irreversible Arrhenius reaction of the first order with the activation
Ea
R
energy
and a frequency factor corresponding to a characteristic time .

3. Effect of Wall Heat Losses in Unobstructed Tubes/Channels


The results have shown that the flame propagation in channels and tubes is affected by the
energy exchange at the surfaces by modifying the thermal expansion ratio, as described in Figure
1. The fuel mixture is preheated by the walls, facilitating burning. This effect is attained in the

beginning of the flame propagation (Fig. 1, left). At a later stage, the fuel mixture flow produced
by the thermal expansion of the burning matter is restricted at the walls, enlarging the flame front
surface and thus accelerating the flame propagation. However, heat gained by the fuel mixture
alters its density, modifying the thermal expansion ratio and, thereby, the flame acceleration
mechanism (Fig. 1, right).

Figure 1. Physical description of the isothermal wall condition coupled to the non-slip wall
friction mechanism of the flame acceleration.
3.1. Wall Temperature Effect
Isothermal walls provide weaker flame acceleration as compared to adiabatic ones. Such a result
has generally been expected since energy is being lost as the flame propagates. However, what is
interesting to notice in the results is the comparison between the isothermal walls at different
temperatures (Fig. 2). While, a priori, one would expect hotter walls to facilitate combustion as
compared to the colder ones, our simulations show the existence of two stages within the flame
spreading scenario: an initial stage, where the expectation is observed indeed; and a second
stage, where the tendency is reversed, namely, the higher the temperature of the isothermal wall,
the slower the flame propagates. The reason behind the establishment of the second trend lays on
the heat transferred to the fresh gas in the radial direction. Specifically, higher wall temperatures
will preheat the unburnt fuel and consequently reduce the thermal expansion, which drives flame
acceleration, thereby reducing the flame front velocity (see Fig. 1). This also explains why larger
flame velocities occur when heat is not lost at the boundaries at all, i.e. in the adiabatic case.

4
3

Dch = 20Lf , = 7

Adiabatic
400 K
600 K
800 K

Rt

1
0
0

4
3

1
0.1

0.2

0.3

0
0

0.4

Dch = 30Lf , = 10

Adiabatic
400 K
600 K
800 K

0.1

0.2

0.3

0.4

Dch = 40Lf , = 10
Adiabatic
400 K
600 K
800 K

Rt

Rt

1
0
0

Adiabatic
400 K
600 K
800 K

Rt

Dch = 20Lf , = 10

0.1

0.2

0.3

0.4

0
0

0.1

0.2

0.3

0.4

Figure 2. Flame tip position along channels with different wall thermal conditions. Plots vary
with channel width and thermal expansion coefficient.
Time evolutions of the scaled flame tip position and its propagation velocity, in the laboratory
reference frame, Fig. 3, show the deceleration observed for warmer walls as compared to colder
ones at the mentioned second stage of flame propagation. The enhanced burning attained by
preheating of the fresh mixture in the beginning of the process is shorty overcome by a reduction
of the thermal expansion.

Ut/SL

10

0.1

400 K
600 K
800 K

10

5
0
0

Dch = 40Lf , = 7

15

400 K
600 K
800 K

Ut/SL

15

Dch = 40Lf , = 7

0.2

0.3

5
0
0

0.4

0.1

0.2

0.3

0.4

Fig. 13. Flame tip position and velocity in channels, varying thermal conditions at the walls.
3.2. Effect of tube/channel width
The changes observed in the flame propagation speed when different widths are considered
diminish in the presence of isothermal walls (Fig. 4). Indeed, a bending of the flame front, and
consequent intensification of its propagation, is more easily attained in narrower channels. Since
the pipe wall is kept at a temperature lower than that of the flame front, burning is prevented in
the vicinity of the walls, thereby reducing the chamber cross section available for the flame front
propagation. However, this positive forcing is not as strong as the mitigation observed when the
thermal expansion is reduced as a result of the heat exchange at the pipe side.
= 8, Ma = 0.014
Adiab.Dch=10Lf

60

300K, Dch=10Lf

U t/SL

80

Adiab.Dch=20Lf

300K, Dch=20Lf

40
20
0
0

0.5

1.5

Figure 4. Flame front tip velocity at adiabatic and isothermal channel wall conditions.
Collapsing of the flame velocities is much more evident at higher wall temperatures, confirming
the dominance of the mitigating effect associated with the reduction of the thermal expansion.
Moreover, while the changes under adiabatic wall conditions are more sensitive to larger overall
thermal expansion coefficients, it is not the case if isothermal walls are in place.

=7

=7

4
3

Rt

Rt

Adiab. Dch=20

Adiab. Dch=20
Adiab. Dch=40

1
0
0

400K, Dch=20
400K, Dch=40

0.1

0.2

0.3

0.4

0
0

800K, Dch=20
800K, Dch=40

0.1

0.2

0.4

= 10

0.3

= 10

Adiab. Dch=30

Rt

Rt

2
Adiab. Dch=30

Adiab. Dch=40

1
0
0

Adiab. Dch=40

400K, Dch=40

0.1

0.2

0.3

Adiab. Dch=40

400K, Dch=30

0
0

0.4

800K, Dch=30
800K, Dch=40

0.1

0.2

0.3

0.4

Figure 5. Evolution of the flame tip position for adiabatic and isothermal channel walls,
including variable channel height. Plots vary with thermal expansion and wall conditions.
3.3. Tubes versus Channels
While the flame propagation in cylindrical axisymmetric tubes resembles that in planar, twodimensional channels, it is also observed the reverse in the velocity trend dominance of colder
isothermal walls over warmer ones after a certain time (Fig. 6). This time interval is noticed to be
longer than that observed in the planar geometry (Fig.7), under the same conditions. Similarly to
the planar case, we can see that the larger initial thermal expansion coefficient (measured at
T fuel mixture 298 K
) generates larger flame speeds, being less sensitive for colder isothermal walls.
8
6

Dtu = 40Lf , = 7

400 K
600 K
800 K

Rt

2
0
0

400 K
600 K
800 K

Rt

Dtu = 40Lf, = 10

2
0.2

0.4

0.6

0.8

0
0

0.2

0.4

0.6

0.8

Figure 6. Flame tip position in tubes, considering different isothermal wall temperatures.

Dtu = Dch = 40Lf, = 7

4
3

Rt

Rt

Adiab. Tube
Adiab. Channel
400 K, Tube
400 K, Channel

1
0
0

Dtu = Dch = 40Lf , = 7

0.1

0.2

0.3

0.4

400 K,
400 K,
800 K,
800 K,

1
0
0

0.1

0.2

0.3

Tube
Channel
Tube
Channel

0.4

Figure 7. Comparison of flame tip propagation in tubes and channels of same width at different
wall thermal conditions.
4. Analysis of Flame Propagation in Obstructed Tubes/Channels
In this section, the analysis is extended to include obstacles in-built at the walls. The resulting
flame spreading process is illustrated in Fig. 8, where a half of the conduit section is shown. As
the flame propagates, delayed combustion, occurring in the pockets between the obstacles,
produces a thermal expansion of the burning mixture inside them, which after adding to that
generated in the core unobstructed region, promotes a jet-flow, generating a powerful flame
acceleration. However, the heat transferred by the walls reduces the thermal expansion ratio,
principally in the pockets, and thereby moderates the acceleration mechanism.

Figure 8. Physical description of the flame propagation in obstructed conduits, including heat
transfer at the surfaces.
4.1. Effect of obstacle size (blockage ratio) on flame propagation
The flame acceleration rates in this configuration is strongly linked to the size of the obstacles,

characterized in our simulations by the blockage ratio , as seen in Fig. 9. The acceleration of
the flame front changes with the obstacles size since it determines the region where the delayed
combustion takes place, which drives this mechanism. In the absence of obstacles at the walls,
the flame accelerates as a result of the frictional forces acting at the walls, although this
mechanism is much weaker.

330

Rt

320

400

=1/9
=1/4
=1/3
=1/2
=2/3

tu

=1/4
=1/2
=2/3

350

Rt

340

D = 30L , = 8

Dch= 24Lf , = 8

310
300
0

0.05

0.1

0.15

300
0

0.05

0.1

0.15

Figure 9. Effect of the obstacles size on the flame tip propagation in adiabatic channels (Left)
and cylinders (Right).
More detailed information can be learned from Fig. 10, where the flame propagation is observed
at different times along the channel. The snapshots vary also with the blockage ratio,
demonstrating the effect it produces on the flame dynamics.

Dch 24 L f
Figure 10. Instantaneous temperature distribution in 2D planar channels,
Dtu 30 L f
axisymmetric cylinders,
(bottom).

(top) and

4.2. Effect of the Surface Temperature


As it can be observed in Fig. 11 (left), the flame propagation attained in adiabatic channels is
approached by isothermal ones when colder temperatures are considered. The preheating of the
fuel-oxidizer mixture ahead of the flame front (and consequent thermal expansion variation) is
negligible if the surfaces are kept at a room temperature (e.g. 298 K). Figure 11 (right) illustrates
the case when the preheating of the fresh mixture is significant. By raising the wall temperature
to 600 K, the thermal expansion is modified by the gained heat, varying the flame propagation
velocity that does not exhibit an exponential trend found in adiabatic and cold wall conditions.

330

330

=1/3
=1/2
=2/3

320

310
300
0

Dch= 20Lf , = 8, Twall= 600K


=1/3
=1/2
=2/3

Rt

Rt

320

Dch= 20Lf , = 8, Twall= 298K

310

0.05

300
0

0.1

0.05

0.1

Figure 11. Flame propagation in channels with variable obstacle size, at different surface
temperatures.
4.2. Effect of the Chamber Width
Bychkov et al. [21] described the flame acceleration attained in obstructed adiabatic conduits as a
Reynolds-independent mechanism, with the Reynolds number defined with respect to the flame
Re Dch / 2 Pr L f
propagation as
. In our results, by keeping the Prandtl number constant, we
have this parameter varying with the channel width only, namely, obstructed channels of width
20 L f 24 L f
30 L f
,
and
will determine the Reynolds number values of 20, 24 and 30, respectively
(Fig. 12). In Figure 12 (left), the Reynolds-independent character of the mechanism is noticed in
cold wall conditions, resulting in a little variation of the flame propagation in channels of
variable widths. However, when the thermal condition at the walls is stressed by raising the wall
temperature (Figure 22; right), an enhanced variation on the flame propagation with the channel
width and, consequently, with the Reynolds number is noticed. In principle, this would obey to
the larger thermal expansion reduction obtained in larger pockets (which dimensions are defined
as a function of the channel width) resulting in slower flame velocities when propagating in
wider channels.
320

= 8, T = 298K, =1/2
wall

320
315

310

310

Rt

Rt

315

Dch=20Lf

305
300
0

= 8, T = 600K, =1/2
wall

Dch=24Lf
Dch=30Lf

0.05

0.1

Dch=20Lf

305
300
0

Dch=24Lf
Dch=30Lf

0.05

0.1

Figure 12. Evolution of the flame tip in channels considering variable width, at different surface
temperatures.

The dominant effect of the obstacle size over the channel width can be observed in Fig. 13. The
finger shape attained by the flame when propagating in these conditions keeps the flame front
away from the channel sides. Having the unobstructed region of the chamber, where the
propagation takes place, determined by the size of the obstacles and considering also that the
thermal energy is mostly transferred by them, we can conclude that the flame dynamics in this
configuration mainly depends on the obstacles geometry and temperature.

320
315

Rt

310

= 8, T = 298K
wall
D ch=20Lf, =1/3
D ch=30Lf, =1/3
D ch=20Lf, =2/3
D ch=30Lf, =2/3

305
300
0

0.05

0.1

Figure 13. Comparison of the channel width and obstacle size effects on the flame dynamics
5. Summary
In this work we have scrutinized the role of thermal boundary conditions at the tube/channel
walls. Specifically, the effect produced by the heat transfer processes at the unobstructed
chamber surfaces has been found to be considerable. Namely, the exponential state of flame
acceleration, the role of the channel/tube width, and the difference between 2D and cylindrical
geometries observed in adiabatic conditions, are mitigated in the presence of isothermal wall
conditions. It is noted that heat transferred to the fresh mixture initially enhances burning, but
then it has been shorty overcome by a consequent reduction of the thermal expansion, mitigating
the flame acceleration. Also, the flame propagation in isothermal tubes and channels approached
that observed under adiabatic conditions when the constant wall temperature was set to be the
room temperature (cold wall condition). With the reduction of the obstacle size (characterized
by the blockage ratio ), the flame acceleration scenario approached that observed in non-slip
unobstructed conditions. In is emphasized, in this respect, that the obstacle size proves to be the
most influential parameter on the flame dynamics, since it determines the pocket region where
the flame acceleration is strengthen by the thermal expansion of the burning matter.
References
[1] Roy G.D. , Frolov S.M. , Borisov A.A., Netzer D.W., Pulse detonation propulsion:
challenges, current status and future perspectives, Prog. Energy Combust. Sci. 30: 545-672
(2004)
[2] Ciccarelli G., Dorofeev S., Flame acceleration and transition to detonation in ducts, Prog.
Energy Combust. Sci. 34: 499-550 (2008)
[3] Dorofeev S., Flame acceleration and explosion safety applications, Proc. Combust. Inst.
33: 2161-2175 (2011)

[4] Kuznetsov M., Alekseev V., Matsukov I., Dorofeev S., DDT in smooth tube filled with
hydrogen-oxygen mixture, Shock Waves 14: 205-215 (2005)
[5] Valiev D., Bychkov V., Akkerman V., Eriksson L.-E., Different stages of flame
acceleration from slow burning to Chapman-Jouguet deflagration, Phys. Rev. E 80: 036317
(2009)
[6] Valiev D., Bychkov V., Akkerman V., C.K. Law, Eriksson L.-E., Flame acceleration in
channels with obstacles in the deflagration-to-detonation transition Combust. Flame 157:
1012-1021 (2010)
[7] Shelkin K.I., Influence of tube roughness on the formation and detonation propagation in
gas, Zh. Eksp. Teor. Fiz. 10: 823-828 (1940)
[8] Zeldovich Ya. B., Barenblatt G.I., Librovich V.B., Makhviladze G.M., Mathematical
Theory of Combustion and Explosion, Consultants Bureau, 1985
[9] Bychkov V., Valiev D., Eriksson L.E., Physical mechanism of ultrafast flame
acceleration, Phys. Rev. Lett. 101: 164501 (2008)
[10] Bychkov V., Petchenko A., Akkerman V., Eriksson L.E., Theory and modelling of
accelerating flames in tubes Phys. Rev. E 72: 046307 (2005)
[11] Akkerman V., Bychkov V., Petchenko A., Eriksson L.E., Accelerating flames in cylindrical
tubes with nonslip at the walls, Combust. Flame 145: 206-219 (2006)
[12] Wu M., Burke M., Son S., Yetter R., Flame acceleration and the transition to detonation of
stoichiometric ethylene/oxygen in microscale tubes, Proc. Combust. Inst. 31: 2429-2436
(2007)
[13] Wu M., Wang C.Y., Reaction propagation modes in millimeter-scale tubes for
ethylene/oxygen mixtures, Proc. Combust. Inst. 33: 2287-2294 (2011)
[14] Bychkov V., Akkerman V., Valiev D., Law C.K., Role of compressibility in moderating
flame acceleration in tubes, Phys. Rev. E 81: 026309 (2010)
[15] Chu R., Clarke J., Lee J.H., Chapman-Jouguet deflagrations, Proc. Roy. Soc. London A
441: 607-623 (1993)
[16] BychkovV., Valiev D., Akkerman V., Law C.K., Gas compression moderates flame
acceleration in deflagration to detonation transition, Combust. Sci. Techn. 184: 1066-1079
(2012)
[17] Valiev D., Bychkov V., Akkerman V., Eriksson L.E., Law C.K., Quasi-steady stages in the
process of premixed flame acceleration in narrow channels, Phys. Fluids 25: 096101
(2013)
[18] Ferguson, C. R., and Keck, J. C., On laminar flame quenching and its application to sparkignition engines Combust. Flame 28:197-205 (1977)
[19] Daou J., Matalon M., Influence of conductive heat losses on the propagation of premixed
flames in channels, Combust. Flame 128: 321-339 (2002)
[20] Demirgok B., Ugarte O., Valiev D., Akkerman V., Effect of thermal expansion on flame
propagation in channels with nonslip walls, Proc. Combust. Inst. (2014), in press.

[21] Bychkov, V.V., Valiev, D. & Eriksson, L.E., 2008. Physical Mechanism of Ultra-Fast Flame
Acceleration. Physical Review Letters, 101, p. 164501.

Você também pode gostar