You are on page 1of 10

Fuel Processing Technology 109 (2013) 179–188

Contents lists available at SciVerse ScienceDirect

Fuel Processing Technology


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

Thermal performance of a micro combustor with heat recirculation


Upendra W. Taywade, Anil A. Deshpande, Sudarshan Kumar ⁎
Department of Aerospace Engineering, Indian Institute of Technology Bombay, Powai, Mumbai, 400076, India

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

Article history: In this paper, numerical and experimental investigations in a three‐step microcombustor have been carried out
Received 7 September 2012 to understand the effect of heat recirculation on flame stabilization behavior with premixed fuel–air mixtures.
Received in revised form 15 October 2012 External heating cup is utilized to enhance the heat recirculation. From numerical simulations, it was observed
Accepted 1 November 2012
that the extent of heat recirculation is a function of cup dimensions, cup material, and flow velocity and mixture
Available online 24 November 2012
equivalence ratio. Heat recirculation has been observed to significantly enhance the flame stability limits and
Keywords:
upper flame stability limits were observed for the range of flow rates investigated during the present work.
Flame stability Stable flames exist for smaller flow rates with a minimum thermal input of ~2.2 W at Φ =0.5. X‐shaped spinning
Heat recirculation flames exist for inter‐mediate flow rate conditions with and without heat recirculation for a broad range of
Micro combustion equivalence ratios. These X‐shaped flames rotate around the axis at ~100–150 Hz frequency. The average com-
Spinning flames bustor wall temperature increases with the flow velocity for the stable flame mode and remain mostly uniform
and well distributed for the X‐shaped spinning flame mode. Heat recirculation helps in increasing the mean wall
temperature of the combustor by ~100–400 K.
© 2012 Elsevier B.V. All rights reserved.

1. Introduction modes at very low and high flow rates and unstable pulsating flame
propagation modes at moderate flow rates. Similar experimental
The processes like hydrogenation, gasification, and pyrolysis of fuels studies on flame dynamics in micro scale channels with different fuel–
require a prescribed temperature distribution and high temperature to air mixtures and wall heat transfer conditions were reported by
achieve higher conversion efficiencies. Due to high surface to volume Ju and Choi [10], Xu and Ju [11], Kumar and co-workers [12–15], Fan
ratio of micro combustion based devices, they have higher heat and et al. [16,17,24], Kariuki and Balachandran [21] and Yang et al. [22].
mass transfer rates and better temperature control. These features of Ronney [6], Veeraragavan and Cadou [7] and Schoegl and Ellzey [8]
the micro-combustion based devices are expected to find applications report the theoretical analysis of the heat recirculation through channel
to make portable, compact, and lightweight fuel processing units walls and their effect on flame propagation in small channels and flame
[1–5]. High power density of the combustion based devices is also stability limits. 2-D analysis of energy and species transport by
expected to result in increased life and reduced weight of small scale Veeraragavan and Cadou [7] reports the effect of heat loss, wall thermal
electronic and microelectromechanical systems (MEMS). The other conductivity and wall geometry on burning velocity and flame extinc-
merits of these systems include lower pollutant emissions, especially tion. Authors have successfully shown that the heat recirculation from
NOx due to lower operating temperatures of these systems as compared post-flame zone to pre-flame zone is the primary mechanism of flame
to the conventional systems [1–4]. The concept of a micro heat engine stabilization in micro channels.
was proposed by Epstein and Senturia [5] in 1997. Since then, signifi- On the application side, Kim et al. [31] have experimentally studied
cant amount of theoretical and experimental works has been reported various configurations of Swiss-roll combustors with propane–air mix-
in the literature towards understanding flame propagation [5–21] at tures and different wall heat transfer conditions. They observed that
small scales and development of small-scale combustion based devices flames can be successfully stabilized for a wide range of mixture equiv-
[22–33]. alence ratios and flow rates by controlling the heat recirculation from
Experimental observations on flame stabilization in a straight quartz combustion products to preheat the incoming fresh mixture through
tube with a wall temperature gradient have been reported by Maruta solid walls. Similar studies on flame stabilization behavior and combus-
et al. [9]. They have reported the formation of stable flame propagation tion performance in other small scale systems such as backward facing
step combustors [22,29,30] and radial micro combustors [14–17,24]
have been reported in the literature. Earlier work [9–18,24–28] mostly
concerns with the flame stabilization in these small scale devices
⁎ Corresponding author. Tel.: +91 22 2576 7124. using heat recirculation through solid walls, altering the flow field,
E-mail address: sudar@aero.iitb.ac.in (S. Kumar). using porous material in the combustor or a catalyst as an aiding

0378-3820/$ – see front matter © 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.fuproc.2012.11.002
180 U.W. Taywade et al. / Fuel Processing Technology 109 (2013) 179–188

mechanism to enhance flame stability limits. X-shaped spinning flames Fig. 1(a) shows the dimensional details of a 3-step backward fac-
in a sudden expansion tube were reported by Kwon et al. [34] for ing microcombustor (model A). The minimum diameter is 2 mm
propane–air and methane–air mixtures. Kurata [35] observed the for- near the inlet and it increases in the subsequent steps. Fig. 1(b)
mation of X-shaped flames in a coaxial burner consisting of two coaxial shows a 3-step backward facing microcombustor (dimensions similar
streams of different premixed mixtures when operated near blowout to model A) covered with an external cup at the exit to enhance heat
conditions. Xu and Ju [11] also observed similar spinning flames in recirculation from hot combustion products to the solid walls of the
meso-scale divergent channels for both lean and rich methane and combustor. To carry out detailed parametric studies, the material of
propane–air mixtures. the outer cup and its dimensions has been changed. A brief summary
The objective of present work is to investigate the combined effect of of the dimensions is shown in Fig. 1b. The experiments were carried
change in the local flow field through backward facing steps and heat out with liquefied petroleum gas (LPG)–air premixed mixtures. The
recirculation through outer solid walls on the flame stability limits. typical composition of LPG gas is ~ 60% butane and ~ 40% propane.
Heat recirculation is enhanced by placing an outer cup around the com- LPG was chosen due to its easy and cheap availability for various do-
bustor and allowing hot combustion products to pass over the combus- mestic and industrial applications.
tor walls. Due to this, the hot combustion products transfer some part of
heat to the combustor walls and subsequently help preheating the in-
coming fuel–air mixture. This increases the average combustor wall 2.2. Details of experimental and computational methods
temperature and results in an improved thermal performance of the
microcombustor. In this paper, the effect of heat recirculation from Fig. 2 shows a schematic diagram of the experimental set-up for the
hot combustion products is delineated and compared with a base case present work. This system consists of fuel and air feed systems, comput-
without any heat recirculation. Parametric studies have been carried erized flow rate measurement and control system and a backward facing
out to understand the effect of heat recirculation on flame position in three step microcombustor (cases A–C). The gas feed system provides a
the micro combustor, flame stability limits, flame propagation modes premixed mixture of fuel (LPG) and dry air from a high pressure storage
and thermal performance of a typical microcombustor. tank. The fuel and air mass flow rates are controlled through a command
module using two precise electric mass flow controllers (AALBORG-GFC)
with 0.5 LPM and 1 LPM flow ranges respectively. Accuracy of the mea-
2. Experimental setup sured mass flow rates is within ±1% of the full scale. To measure the wall
temperature profiles of the microcombustor, three K-type thermocou-
2.1. Micro-combustor configurations ples (0.5 mm diameter) were arranged on the outer wall of the combus-
tor. These thermocouples were arranged at the junctions of three steps of
Fig. 1 shows the dimensional details of the microcombustor con- the combustor as shown in Fig. 3. In case of model B, a similar procedure
figurations employed for numerical and experimental studies. These was adopted and the thermocouples were placed on the outer wall of the
microcombustors were designed based on the continuation of the cup. For certain cases, a thermocouple was inserted inside the cup to
previous work [29,33] on two and three-step microcombustors. The measure the outer wall temperature of the combustor. Thermal contact
purpose of using backward facing steps is to provide a sudden expan- resistance between the thermocouple and combustor, heat loss due to
sion to the incoming mixture which alters the velocity profile near conduction, convection and radiation through thermocouple junction
the wall and helps in enhancing the flame stability limits in such were estimated. Thermal contact resistance was estimated from the rela-
microcombustors. tion given in ref. [36]. It was found that measured temperature error due

Fig. 1. Configuration and dimension details (a) model A without any heat recirculation cup (b) models B and C with heat recirculation cup.
U.W. Taywade et al. / Fuel Processing Technology 109 (2013) 179–188 181

Fig. 2. Schematic diagram of the experimental setup.

to heat loss is less than 3% for temperatures below 1000 K. Hence, the to complete one simulation. Heat losses from solid walls were considered
corrections to the measured temperature were not included in the and assumed to be 15.0 W/m2K and thermal radiation was neglected
present paper. [29]. Velocity inflow condition was employed at the inlet and pressure
The numerical modeling has been carried out to qualitatively under- outlet boundary condition at the exit of the combustor. Chemical reaction
stand the role of heat recirculation on flame propagation and flame stabil- was modeled with a simplified single-step reaction mechanism to under-
ity limits in a backward-step combustor. Therefore, a simple 2-D model stand the formation of products such as CO2 and H2O from the combus-
has been chosen along with conjugate heat-transfer and single step reac- tion zone [38]. Numerical simulations were carried out for the same
tion to model the combustion process. Steady-state analysis has been car- operating conditions as that of experiments. Initial computations were
ried to obtain qualitative results on flame stability and wall temperature started with 10,000 grid points and grid adaption was employed to appro-
profiles to help explain the present experimental results. Numerical sim- priately resolve the reaction zone structure within the flame front and the
ulations were carried out using a general purpose CFD code Fluent 6.3.26 results obtained are grid independent because the minimum grid size
[37] and the results obtained were analyzed along with the experimental maintained within the reaction zone of the propagating flame front was
results for understanding the effect of heat recirculation on flame stabili- as small as 10–20 μm [38].
zation behavior in these microcombustors. Two-dimensional Navier–
Stokes equations were solved in cylindrical coordinates along with 3. Result and discussion
energy and species conservation equations. The flow Reynolds number
varies in the range of 150–800 for the given flow rate conditions. There- 3.1. Preliminary observations
fore, laminar flow calculations were carried out to obtain the distributions
of velocity, pressure, species and temperature in the computational do- During the preliminary investigations on flame stabilization, the char-
main. The solution was considered converged when scaled residuals of acteristics of the combustion exhibited by the premixed LPG–air mixture
mass, momentum, energy and species had dropped by five orders of mag- were studied for its stable modes inside the micro-combustor. Varied
nitude and there was no appreciable change in the respective residuals flow rates and mixture equivalence ratios were maintained. It was ob-
further. It took approximately 48 h of CPU time on a desktop computer served that the flame inside the micro-combustor changes its shape

Fig. 3. Schematic of the thermocouple arrangement on the walls of the microcombustor.


182 U.W. Taywade et al. / Fuel Processing Technology 109 (2013) 179–188

and position with a change in the flow rate and mixture equivalence combustor with different conditions of heat recirculation (cases A–C)
ratio. Direct photographs of various combustion modes exhibited for a for stoichiometric methane–air mixtures at a mixture flow velocity of
range of mixture flow rates and mixture equivalence ratios are shown 1.0 m/s. Top part of Fig. 5 shows the temperature contours and bottom
in Fig. 4. For lean mixtures and lower flow rates, the flame shape remains part shows the reaction rate contours along with flow streamlines in the
mostly C-shaped as shown in Fig. 4(a). The flame is stabilized mostly in microcombustors. It is clear from Fig. 5a that for case A, the flame is sta-
the first and second steps of the combustors (for all combustors). How- bilized at the start of the third step of the microcombustor. For cases B
ever, as the inlet velocity is increased, the effect of heat re-circulation on and C, the flame is stabilized at the start and middle of the second
flame shape can be seen clearly in the subsequent frames. step respectively as shown in Fig. 5b and c. Flame stabilization point
For model A, at high flow velocities, an X-shaped flame is observed moves upstream due to increased heat recirculation from hot combus-
as shown in Fig. 4(c). A detailed study of these flames with a tion products. The comparison of flame stabilization point shows that
high-speed camera has revealed that it is a thin reaction front rotating the heat recirculation is maximum for case B followed by case C and
at a very high frequency in the range of 80 to 150 Hz due to which it ap- case A. The extent of preheating of incoming fuel–air can be determined
pears to be X-shaped to the naked eye [11]. This propagating flame front from the location of 600 K temperature contour shown in Fig. 5a–c. It is
is mostly anchored at the step of the channel. It was observed that a clear from these figures that the 600 K contour reaches the start of the
sudden change in the velocity of the mixture or the equivalence ratio first step for case B indicating maximum heat recirculation as compared
changed the flame shape to conical shaped flame momentarily and to cases C and A respectively. For case A, heat recirculation occurs only
soon reverted back to original X-shaped flame. This mode appeared through solid walls of the combustor. Therefore, 600 K temperature
for flow velocities higher than 1.5 m/s for very rich mixtures and at contour is limited to a very small distance in the upstream direction.
lower velocities (b 1.5 m/s) for lean mixtures. A part of the unburnt Fig. 6 shows the temperature distribution of the combustor wall
mixture burns at the exit plane due to lack of sufficient oxidizer in the obtained from numerical results for these different cases. It is clear
mixture as shown in Fig. 4(b). from this figure that with an increase in the heat recirculation from
For model B, with heat recirculation, the appearance of the various the hot combustion products for case B and case C, the wall tempera-
flame propagation modes is shown in Fig. 4(d–f). At Ф = 0.8 and a tures are relatively higher than for a simple combustor case A. The
flow velocity of 1.3 m/s, a flat flame is observed as shown in Fig. 4(d) peak in the wall temperature profile also indicates the approximate
due to heat recirculation than compared to a conical flame for same position of the flame stabilization point. Therefore, it is expected
condition without heat recirculation (not shown here). For Ф = 1.0, a that heat recirculation will significantly enhance the flame stability
conical flame anchors at the third step and stabilizes as clear from limits resulting in increased flame stability limits for cases B and C
Fig. 4(e). This type of flame is generally very stable, remains anchored compared to case A.
to third step and seen for velocities ranging from 1.5 m/s to 9.0 m/s
for 0.8 b Ф b 1.1. However, for very lean and rich mixtures and higher ve- 3.3. Flame stability limits
locities, X-shaped flames were observed as shown in Fig. 4(f).
It has been observed that a self-sustained flame can be achieved only
3.2. Flow field and related details for a range of mixture velocities and equivalence ratios. Fig. 7 shows the
flame stability limits of a 3-step quartz micro-combustor with premixed
Fig. 5 shows the detailed distribution of temperature, reaction rate, LPG-air mixtures for various microcombustors (models A–C). A stable
and flow streamlines obtained from numerical modeling of present flame can be obtained for mixtures as lean as Ф = 0.5. The flame

Fig. 4. Flame propagation modes in the combustor (a–d) for model A (without heat re-circulation) (e–h) for model B (with heat re-circulation).
U.W. Taywade et al. / Fuel Processing Technology 109 (2013) 179–188 183

Fig. 5. Numerical predictions of the flame stabilization in the combustor for stoichiometric methane–air mixtures at different heat recirculation conditions: (a) case A (b) case B and
(c) case C. Temperature contours shown in top part and reaction rate contours along with flow streamlines (dashed lines) on the bottom part of the figure.

stability limits of a three step combustor are relatively much wider as 150 SSCM in a two-step combustor. Due to these advantages, a
compared to a two-step combustor [29]. This is perhaps due to the rea- three-step combustor has been considered for detailed investigations
son that increase in the number of steps helps in gradual flow deceler- in the present work. The flame stability limits for the case of heat
ation and enhances the flame stability limits. A stable flame is obtained recirculation (cases B and C) are much wider than for the case of no
for a minimum thermal input of approximately 3.5 W in a three-step heat recirculation. For instance, a stable flame can be obtained for mix-
combustor without any heat recirculation (case A). This limit of mini- ture velocities as small as 0.4 m/s and these flames are stabilized in the
mum thermal input is much smaller than a two-step combustor first step of the combustor. This corresponds to a minimum thermal
(~5 W) [29]. At lower limits, a stable flame can be obtained for a mix- input of ~2.2 W which is relatively much smaller as compared to the
ture flow rate of 95 SCCM in a three-step combustor as compared to minimum thermal input achieved for two step combustors, ~5 W
184 U.W. Taywade et al. / Fuel Processing Technology 109 (2013) 179–188

1200 the flame remains stabilized either in the third step or near the exit of
Case A combustor and the effect of heat recirculation is enhanced due to higher
Case B overall heat release. This helps in preventing the flame blow off even for
Wall temperature (K)

Case C very high flow rate conditions. The qualitative comparison of numerical
900 and experimental results can be observed here. Since the heat
recirculation and solid wall temperature (Fig. 6) are maximum for
case B than cases C and A, no upper flame stability limits were observed
to exist for case B combustor. For case C, flame blow off was observed to
occur at Ф = 0.5 and 0.6 and for case A, upper flame stability limits are
600
observed to exist for the complete range of mixture equivalence ratios.
Fig. 8(a) shows the regimes of existence for X-shaped spinning
flame mode (shown in Fig. 4c) for model A (without heat
re-circulation). A spinning flame exists for a range of mixture velocities
300 and equivalence ratios. The spinning flame limits are narrow for leaner
0 0.01 0.02 0.03 0.04 0.05 0.06
mixtures and becomes wider for rich mixtures as compared to models B
Wall length (m) and C (shown in Fig. 8b), where two separate wider spinning flame
Fig. 6. Numerical predictions of the wall temperature profile for different combustor
limits are observed to exist for lean and rich mixtures. Fig. 8(b) shows
cases. the limits of formation of the spinning flame for models B and C. It is in-
teresting to note that the spinning flame modes exist for a wide range of

[29]. The lower flame stability limits remain almost constant for Ф b 1.25
and then increase almost linearly up to Ф = 1.5 for both models B and C.
For model A, the upper flame stability limits were observed to increase
with mixture equivalence ratio. For flow velocities higher than the
upper flame stability limits, the flame blow off occurs. For case B, exper-
iments were carried out for flow velocities up to 9 m/s and no flame
blow off was observed for all the mixture equivalence ratio conditions.
Similarly for model C, the lower limits are almost similar to that of
model B. Upper stability limits were observed to exist for case C at
Ф = 0.5 and 0.6 as compared to other equivalence ratios. For higher
mixture equivalence ratios, no flame blow off was observed till 9 m/s
flow velocity. Overall, it is interesting to note the effect of heat
recirculation at lower flame stability limits as it is clear from these
three cases that the minimum thermal input can be significantly re-
duced to as small as 2.2 W as compared to ~3.5 W for case A without
any heat recirculation. This indicates that heat recirculation from hot
combustion products helps in increasing both upper and lower flame
stability limits significantly for these microcombustors. This is due to
the fact that for very small flow rates, the flame is mostly stabilized in
the first step and flame quenching is governed by heat loss. Any small
contribution through heat recirculation from hot combustion products
helps in improving the lower flame stability limits. At higher velocities,

Fig. 8. (a) Spinning flame limits without heat recirculation, model A. (b) Spinning
Fig. 7. Flammability limits of the combustors with and without heat re-circulation. flame limits with quartz heat recirculation cup, models B and C.
U.W. Taywade et al. / Fuel Processing Technology 109 (2013) 179–188 185

equivalence ratios and mixture velocities. For equivalence ratio range of spins at a very high frequency and hence appears like X-shaped to the
0.8 ≤ Ф ≤ 1.0 (model B) and 0.9 ≤Ф ≤ 1.1 (model C) and at a flow veloc- naked eyes. It also shows that the spinning flame front is anchored at
ity of ~1.0 m/s, X-shaped flame suddenly appears in the second step. the backward step of the channel. Fig. 9(b) shows the variation in the
For this mixture equivalence ratio range, X-shaped spinning flame is frequency of spinning flames at different equivalence ratios for a mix-
formed only for very small range of mixture velocities. An increase in ture velocity of V= 2 m/s for LPG–air mixture. It is observed that fre-
the mixture velocity results in the flame stabilization in the third step quency increases with an increase in equivalence ratio till Ф = 1.0 and
as a conical flame. For other equivalence ratios, the spinning flame is then decreases for rich mixtures.
stabilized at the third step. These present experiments on various com- Further investigations were carried out to understand the effect of
bustors (cases A–C) have clearly helped in delineating the role of heat Lewis number on the formation of X-shaped spinning flame, particular
recirculation/loss on the formation of these X-shaped flames. Heat for the case of heat recirculation. The effective Lewis number was esti-
recirculation (cases B and C) significantly reduces the formation of mated at 300 K and 500 K conditions for a range of mixture equivalence
these flames as clear from Fig. 8a and b. Fig. 9(a) shows the direct ratios of propane–air mixture to see the effect of indirect heat
images of the spinning flames captured with a high speed camera recirculation from flame to wall and then from wall to fresh fuel–air
(Phantom, ver. 12.1) at a speed of 2000 fps. The figure clearly shows mixture. It is observed that effective Lewis number decreases with in-
that the X-shaped flame is a thin reaction zone of the flame which crease in temperature and this is expected to result in the reduced

Fig. 9. (a) Time sequence of spinning flame for LPG–air mixture at Ф=0.9 and V=2.0 m/s (time shown in each frame is in ms) (b) frequencies of spinning flames for various equivalence
ratios at V=2.0 m/s for LPG–air mixture.
186 U.W. Taywade et al. / Fuel Processing Technology 109 (2013) 179–188

thermal flame wall coupling. These X-shaped spinning flames appear combustor for a range of mixture flow velocities. The flame position
due to flame bifurcations for a wide range of mixture Lewis numbers shifts towards upstream with heat recirculation. This is clear from
[6]. The appearance of these spinning flames can be attributed to strong the curves corresponding to the case of heat recirculation and with-
thermal-wall coupling [6,9]. The transition between a spinning and out heat recirculation at Ф = 0.8. An increase in the mixture equiva-
steady flame mode can be further controlled through external cooling lence ratio also results in the flame stabilization at an upstream
(case A) or heating of the microcombustor walls (case B and case C), location as clear from the curves corresponding to Ф = 0.6 and 0.8.
thus confirming the fact that flame-wall coupling leads to the appear- For the case of no heat recirculation at Ф = 0.8, the flame initially sta-
ance of these spinning flames. bilizes in the downstream region of first step and moves further
downstream with an increase in the flow velocity. The stabilized
flame blows-off for flow velocities higher than ~ 3 m/s. There is very
3.4. Variation of flame position with mixture flow velocity little change in the flame position due to a change in flow velocity
(1.0 m/s to 1.2 m/s). This indicates that the flame stability is signifi-
Fig. 10(a) shows the variation of the flame tip position with mix- cantly enhanced by the backward step of the combustor [22,29]. For
ture flow velocity for two cases of lean mixture conditions, Ф = 0.6 a case with heat recirculation, the flame stability limits are much
and 0.8. This comparison has been carried out to further delineate wider and flame is stabilized even at the exit of the combustor for
the effect of heat recirculation and mixture equivalence ratio on higher flow velocities and flame does not blow off even for velocities
flame position in the combustors. It can be seen from Fig. 10(a) that as high as 9 m/s. It can be further observed from Fig. 10(a) that at
a stable flame is obtained for heat recirculation case even at Ф = 0.6. each step, there is a change in the gradient of the curve representing
The flame is stabilized in the first, second and third step of the the flame position. This indicates that the flame stability is enhanced
when the flame is anchored at the backward facing step and the
change in flame position is very small with an increase in mixture ve-
locity. Fig. 10(b) shows the variation of the flame length with mixture
velocity for different operating conditions. It is clear from Fig. 10(b)
that for heat recirculation case, length of the flame decreases with
an increase in equivalence ratio. This is due to the fact that mixture
burning velocity increases with Ф. For Ф = 0.8 and heat recirculation
case, a conical flame is anchored at the third step for all the mixture
velocity conditions and the length of this flame increases almost line-
arly with an increase in the mixture velocity(represented by curve C).
For Ф = 0.6, the conical flame is anchored at exit plane for flow veloc-
ities greater than 2.7 m/s (curve C1). For no heat recirculation case,
X-shaped flame is observed at Ф = 0.8 for a velocity range of 1.5 to
2.3 m/s (curve X). At Ф = 0.6, for heat recirculation case, X-shaped
flame is observed to exist for a velocity range of 1.5 to 2.6 m/s (curves
X and X1).
Fig. 11 shows the variation of the flame position with mixture equiv-
alence ratio for a range of flow velocities. A dip is observed for stoichio-
metric mixtures and slightly rich mixtures. This is due to the fact that
the maximum mixture burning velocity is achieved for stoichiometric
and slightly rich mixtures. Therefore, flame moves upstream with an in-
crease in mixture equivalence ratio till Ф = 1.1 and then starts moving
downstream again. This is quite consistent for a large range of mixture
velocities as shown in Fig. 11.

Fig. 10. (a) Variation of flame position with mixture velocity for models A and B with
quartz cup (b) variation of flame length with mixture velocity for models A and B with Fig. 11. Flame propagation varying with mixture equivalence ratio for different flow
quartz cup. velocities with quartz heat re-circulation cup for model B.
U.W. Taywade et al. / Fuel Processing Technology 109 (2013) 179–188 187

Fig. 12. Wall temperature profiles along the length of combustor without heat
re-circulation.

Fig. 14. Temperature variation with inlet velocity for quartz and aluminum cup for
model B.
3.5. Wall temperature profiles

To study the effect of flow velocity, mixture equivalence ratio and 3.5.2. Effect of mixture equivalence ratio on outer wall temperature
heat recirculation on the wall temperature of the micro-combustor, Fig. 13 shows the variation of the average wall temperature for a
experiments were carried out for a range of mixture velocities and range of mixture equivalence ratios and flow velocities. The details
equivalence ratios. The results are presented below. of the position of the thermocouples are shown in Fig. 3. It can be ob-
served that maximum wall temperature is obtained for stoichiometric
or slightly rich mixtures. An increase in the flow velocity increases the
3.5.1. Effect of inlet velocity on outer wall temperature
wall combustor wall temperature. The average wall temperature of
The variation of the average outer wall temperature near the flame
the combustor can be varied from 400 to 700 K depending on the
location for different flow velocities is shown in Fig. 12. The details of
mixture flow velocity and equivalence ratio.
the position of the thermocouples are discussed in Fig. 3. The values of
A comparison of wall temperatures for an aluminum cup and a quartz
the temperatures plotted in this figure are the average temperature
cup with model B configuration is shown in Fig. 14. The outer wall
values recorded by these three thermocouples. The experimentally
temperature is measured at the middle of the third step of the combus-
measured wall temperature for case A varies in the range of 400–
tor. The outer wall temperature increases with an increase in the flow ve-
600 K. It is interesting to note that a stable C-shaped flame is formed
locity. The wall temperature for a quartz cup is ~200 K higher than for
for lower velocities and the average wall temperature increases with
aluminum cup when mixture velocities are higher than 2.5 m/s. The dif-
flow velocity. For velocities in the range of 1.6 to 4.0 m/s, X-shaped
ference can be attributed to higher thermal conductivity of aluminum
flame appears and the variation of the wall temperature is relatively
material than quartz material. Due to higher thermal conductivity, heat
small. A more uniform temperature profile can be obtained during the
losses from aluminum cup are much higher resulting in lower wall tem-
operation of the combustor in X-shaped flame propagation mode be-
peratures. It is further to be noted that for higher flow velocities, the
cause of its spinning nature. For higher velocities (>4 m/s) conical
measured temperatures at the combustor walls were as high as 1200 K
flame is formed at the exit plane of combustor resulting in a decrease
for stoichiometric fuel–air mixtures. This behavior can be attributed to
in the wall temperature.
the excess heat of the exhaust gases recirculated through the solid
walls resulting in excess enthalpy combustion and higher wall
temperatures.

4. Conclusions

Numerical and experimental studies on flame stabilization charac-


teristics have been carried out with premixed LPG–air mixtures for
two different cases; with and without heat recirculation in a three
step backward facing cylindrical combustor. The effects of heat
recirculation, wall thermal conductivity, flow velocity and mixture
equivalence ratio on flame stability limits, wall temperature profiles
and flame propagation modes have been studied. Numerical studies
helped in understanding the flame stabilization behavior in different
combustors and effect of heat recirculation on flame stability limits
and wall temperature profiles. Heat recirculation helps in enhancing
the flame stability limits significantly. No upper flame stability limits
were observed for models B and C as compared to model A. The lower
flame stability limits of models B and C are significantly enhanced
resulting in a minimum thermal input as small as 2.2 W. Regimes of
spinning modes are observed to exist for models B and C which are sig-
Fig. 13. Variation of outside wall temperature of the quartz cup for model B. nificantly different from model A due to increased heat recirculation
188 U.W. Taywade et al. / Fuel Processing Technology 109 (2013) 179–188

from hot combustion products. The formation of these spinning and [17] A.W. Fan, S. Minaev, S. Kumar, W. Liu, K. Maruta, Combustion and Flame 153
(2008) 479–489.
stable combustion modes depends on the extent of heat recirculation, [18] C.J. Evans, D.C. Kyritsis, Combustion Science and Technology 183 (2011) 847–867.
flow rate and mixture equivalence ratios. The average combustor wall [19] S. Heatwole, A. Veeraragavan, C.P. Cadou, S.G. Buckley, Nanoscale and Microscale
temperature with quartz cup is ~100–200 K higher than aluminum Thermophysical Engineering 13 (2009) 54–76.
[20] C.J. Evans, D.C. Kyritsis, Proceedings of the Combustion Institute 32 (2009)
cup depending on mixture velocities. Outer wall temperature in the 3107–3114.
range of 500–900 K can be obtained depending on application require- [21] J. Kariuki, R. Balachandran, Experimental Thermal and Fluid Science 34 (2010)
ments. These characteristics of wide flame stability limits, uniform wall 330–337.
[22] W.M. Yang, S.K. Chou, C. Shu, Z.W. Li, H. Xue, Applied Thermal Engineering 22
temperature profile for a range of flow and equivalence ratio conditions (2002) 1777–1787.
make these combustors a good choice for various heating and power [23] K.H. Lee, O.C. Kwon, Combustion and Flame 153 (2008) 161–172.
producing applications from combustion based devices at small scales. [24] A.W. Fan, S. Minaev, E. Sereshchenko, R. Fursenko, S. Kumar, W. Liu, K. Maruta,
Proceedings of the Combustion Institute 32 (2009) 3059–3066.
[25] Y. Wang, Z. Zhou, W. Yang, J. Zhou, J. Liu, Z. Wang, K. Cen, Experimental Thermal
and Fluid Science 35 (2011) 1451–1457.
References [26] G.P.S. Sahota, B. Khandelwal, S. Kumar, Energy Conversion and Management 52
(2011) 3206–3213.
[1] A.C. Fernandez-pello, Proceedings of the Combustion Institute 29 (2002) [27] Y. Wang, Z. Zhou, W. Yang, J. Zhou, J. Liu, Z. Wang, K. Cen, Energy Conversion and
883–899. Management 51 (2010) 1127–1133.
[2] Y. Ju, K. Maruta, Progress in Energy and Combustion Science 37 (2011) 669–715. [28] L.Q. Jiang, D.Q. Zhao, X.H. Wang, W.B. Yang, Energy Conversion and Management
[3] S.A. Lloyd, F.J. Weinberg, Nature 251 (1974) 47–49. 50 (2009) 1308–1313.
[4] N.S. Kaisare, D.G. Vlachos, Progress in Energy and Combustion Science 38 (2012) [29] B. Khandelwal, G.P.S. Sahota, S. Kumar, Journal of Micromechanics and
321–359. Microengineering 20 (2010) 095030.
[5] A.H. Epstein, S.D. Senturia, Science 276 (1997) 1211. [30] J. Li, S.K. Chou, G. Huang, W.M. Yang, Z.W. Li, Experimental Thermal and Fluid Science
[6] P.D. Ronney, Combustion and Flame 135 (2003) 421–439. 33 (2009) 764–773.
[7] A. Veeraragavan, C.P. Cadou, Combustion and Flame 158 (2011) 2178–2187. [31] N.I. Kim, S. Kato, T. Kataoka, T. Yokomori, S. Maruyama, T. Fujimori, K. Maruta,
[8] I. Schoegl, J.L. Ellzey, Combustion and Flame 151 (2007) 142–159. Combustion and Flame 141 (2005) 229–240.
[9] K. Maruta, T. Kataoka, N.I. Kim, S. Minaev, R. Fursenko, Proceedings of the [32] S. Kumar, K. Maruta, S. Minaev, Journal of Micromechanics and Microengineering
Combustion Institute 30 (2005) 2429–2436. 17 (2007) 900–908.
[10] Y. Ju, C.W. Choi, Combustion and Flame 133 (2003) 483–493. [33] A.A. Deshpande, S. Kumar, Applied Thermal Engineering 51 (2013) 91–101.
[11] B. Xu, Y. Ju, Proceedings of the Combustion Institute 31 (2007) 3285–3292. [34] M.J. Kwon, B.J. Lee, S.H. Chung, Combustion and Flame 105 (1996) 180–188.
[12] B. Khandelwal, S. Kumar, Applied Thermal Engineering 30 (2010) 2718–2723. [35] O. Kurata, Combustion and Flame 152 (2008) 206–217.
[13] M. Akram, S. Kumar, Combustion and Flame 158 (2011) 915–924. [36] J.P. Holman, S. Bhattacharyya, Heat Transfer, ninth edition Tata McGraw Hill, New
[14] S. Kumar, K. Maruta, S. Minaev, R. Fursesnko, Physics of Fluids 20 (2008) 024101. Delhi, 2008.
[15] S. Kumar, K. Maruta, S. Minaev, Physical Review E 75 (2007) 016208. [37] Fluent 6 User Guide, Lebanon, NH, 2003.
[16] A.W. Fan, S. Minaev, S. Kumar, W. Liu, K. Maruta, Journal of Micromechanics and [38] S. Kumar, Combustion Science and Technology 183 (8) (2011) 779–801.
Microengineering 17 (2007) 2398–2406.

You might also like