Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Combustion and Flame 138 (2004) 97–107

www.elsevier.com/locate/jnlabr/cnf

A CFD study of propane/air microflame stability


D.G. Norton and D.G. Vlachos ∗
Department of Chemical Engineering and Center for Catalytic Science and Technology (CCST), University of Delaware,
Newark, DE 19716-3110, USA
Received 7 October 2003; received in revised form 26 March 2004; accepted 5 April 2004
Available online 8 May 2004

Abstract
A two-dimensional elliptic computational fluid dynamics model of a microburner is solved to study the effects
of microburner wall conductivity, external heat losses, burner dimensions, and operating conditions on combus-
tion characteristics and the steady-state, self-sustained flame stability of propane/air mixtures. Large gradients are
observed, despite the small scales of the microburners. It is found that the wall thermal conductivity is vital in
determining the flame stability of the system, as the walls are responsible for the majority of the upstream heat
transfer as well as the external heat losses. Furthermore, there exists a range of flow velocities that allow stabilized
combustion in microburners. It is found that the microburner dimensions strongly affect thermal stability. Engi-
neering maps denoting flame stability are constructed and design recommendations are made. Finally, comparisons
with methane/air systems are made.
 2004 The Combustion Institute. Published by Elsevier Inc. All rights reserved.

Keywords: Propane; Microburners; Computational fluid dynamics; Flame stability; Extinction; Thermal management

1. Introduction the production of hydrogen for fuel cell applica-


tions [3].
Unfortunately, the benefits arising at the mi-
Microburners may play a vital role in the portable
croscale are overshadowed by major difficulties in
production of energy. Hydrocarbons have an energy creating working microburners. In 1817, Davy per-
density significantly higher than that of the traditional formed experiments showing the inability of flames
Li batteries (40 vs 0.5 MJ/kg) that are currently used to propagate between gaps of submillimeter scale [4].
in laptops, cellular phones, and other portable elec- His work was followed by many other groups, whose
tronics devices [1]. The small scales in microburners work confirmed that depending on geometry, com-
result in lower combustion temperatures due to en- position, and flow rate, hydrocarbon/air flames are
hanced heat-transfer coefficients. Thus, we propose typically quenched when confined within spaces with
that microburners could possibly reduce the gas-phase critical dimensions < 1–2 mm [5–9]. The two pri-
production of NO [2]. Finally, microburners can also mary mechanisms for quenching in these systems are
serve as efficient sources of heat for endothermic re- thermal and radical quenching [2,10,11]. Increased
actions, such as steam reforming and ammonia de- heat-transfer coefficients are inherent to microscales,
composition, in integrated microchemical systems for because for a fixed Nusselt number, the heat-transfer
coefficient scales with the inverse of the length scale.
The high heat-transfer rates increase the heat lost
* Corresponding author. Fax: (302)-831-1048. from the reaction, reducing the operating tempera-
E-mail address: vlachos@che.udel.edu tures and causing the combustion to extinguish. At
(D.G. Vlachos). the same time, the increased mass transfer within

0010-2180/$ – see front matter  2004 The Combustion Institute. Published by Elsevier Inc. All rights reserved.
doi:10.1016/j.combustflame.2004.04.004
98 D.G. Norton, D.G. Vlachos / Combustion and Flame 138 (2004) 97–107

the system, coupled with the high surface-area-to- tive agreement with our simulations [13]. However,
volume ratio, causes radical adsorption onto the walls, it is unclear how much of the TSB experimental re-
followed by radical recombination. This dearth of sults transfer to the microscale, where the flow is
radicals quenches the homogeneous chemistry [12]. severely confined. Another idea for stabilizing com-
Another mechanism for loss of stability is blowout, bustion within mesoscale (∼ 1 mm) and macroscale
which occurs when the burner exit velocity exceeds ( 1 mm) tubes revolves around the work of Lloyd
the flame burning velocity [7]. In this case, the reac- and Weinberg on heat-recirculating burners, where
tion front shifts downstream with increasing velocity hot combustion products are used to preheat the in-
and eventually exits the burner. The competition be- coming feed [18,19]. Ronney and co-workers have
tween the shifting of the reaction zone and thermal extended this idea on mesoscale “Swiss roll” heat-
quenching has been observed in elliptic models of recirculating burners, with gap sizes on the order of
microscale systems for methane [13] and mesoscale 3 mm, for the homogeneous and heterogeneous com-
systems for propane [14]. bustion of propane in air [1,20]. In recent work, Ron-
Recent experiments have demonstrated that it is ney modeled a countercurrent heat-recirculating com-
feasible to stabilize homogeneous methane/oxygen bustor [21]. The system consisted of a cold reactant
flames between parallel plates with gaps smaller than inlet that fed into a well-stirred reactor (WSR). The
1 mm [15,16]. This is accomplished by modifying products from the WSR then flowed countercurren-
the surface to make it chemically inactive, to elimi- twise past the incoming feed to preheat it. The hot
nate radical quenching, and insulating the burner, to and cold tubes were modeled by a one-dimensional
reduce thermal quenching. The chemical inactivation model, with convective heat-transfer coefficients to
process involves high-temperature annealing to heal model the heat transfer between them and the exterior.
crystal defects and surface cleaning with deionized The reaction was assumed to take place within the
water, hydrochloric acid, and hydrogen peroxide to WSR. It was shown that the axial conduction within
remove ionic and heavy metal contaminants. the wall had a major effect on the operating limits of
In our recent computational fluid dynamics (CFD) the combustor. Even a small wall thermal conductiv-
work [13], overall heat management was shown to ity results in significantly higher required flow rates to
play a critical role in determining homogeneous flame maintain stabilized combustion, whereas when the ax-
stability of methane/air mixtures in microburners. The ial wall conduction is ignored, the minimum required
thermal conductivity of the wall allowed both the vi- flow rate disappears. Ronney’s analysis emphasizes
tal upstream heat transfer for preheating the feed to the importance of heat transfer as our CFD modeling
the ignition temperature and detrimental heat losses did [13].
to the exterior. When the thermal conductivity is too Despite previous work, there are a number of an-
low, the upstream heat transfer through the walls is swered questions regarding flame stability at the mi-
choked, and the system blows out. At the other ex- croscale. Examples include the roles of fuel and mi-
treme of high thermal conductivity, the wall and fluid croburner dimensions in flame stability. In this work
temperature profiles flatten, causing delocalized reac- we extend our previous modeling work to the steady-
tion fronts and an increased external hot area for heat state self-sustained microcombustion of propane/air.
losses. As a result extinction is observed. Two-dimensional (2D), fully elliptic simulations are
The significant role of upstream heating in sta- performed that explicitly treat heat and mass trans-
bilizing combustion in a tube was in fact realized fer in the fluid as well as heat transfer in the wall.
several years ago by Churchill [17], who developed The effects of wall thermal conductivity, external heat
thermally stabilized burners (TSB). These TSB con- losses, operating conditions, and microburner dimen-
sisted of ceramic cylindrical tubes > 7 mm in diam- sions on flame stability and combustion characteris-
eter into which fuel and air were fed. Flame fronts tics are discussed. In order to understand how dif-
were stabilized within the tubes by thermal feedback ferent hydrocarbons behave, the differences between
caused by conduction through the walls and radia- propane/air and methane/air systems are also investi-
tion between the walls to preheat the incoming feed. gated.
To maximize the heat transfer between the fluid and
the walls, the flow rates were such that the flow was
turbulent upstream of the combustion. In the reac- 2. Model
tion zone further downstream, the viscosity increased
due to the increased temperature, resulting in lami- The burner is modeled as two parallel plates that
nar flow. These systems were found to be very sta- are infinitely wide, 1 cm long, and distance L apart.
ble to minor disturbances, but the range of flow rates The plate thickness is Lw . For most simulations L =
that allowed stabilized combustion was very limited. 600 µm and Lw = 200 µm (unless otherwise noted).
These experimental findings are in general qualita- Premixed, nonpreheated propane/air mixtures are fed
D.G. Norton, D.G. Vlachos / Combustion and Flame 138 (2004) 97–107 99

the fluid and the solid are calculated explicitly with


the 2D elliptic models without any further simplifica-
tions. The exterior convective heat-transfer coefficient
is only used for the calculation of the heat flux of
the exterior wall edge boundary condition. This heat-
transfer coefficient lumps the details of heat loss from
the microburner and of the process that utilizes the
heat generated by the burner. The left and right wall
Fig. 1. Schematic of the computational domain (not to scale edges are taken to be insulated (zero flux boundary
for ease of visualization).
condition).
Nonuniform node spacing is employed in this
to the inlet of the microburner, and hot product gases work, with more nodes in the reaction zone. The num-
exit the microburner. Due to the aspect ratio, the sys- ber of nodes varies depending on dimensions, but
tem is simplified to a 2D one and the plane of symme- the simplest one consists of 120 axial nodes by 60
try between the two plates allows simulations of only transverse nodes, totaling approximately 7200 nodes.
half of the microburner. A schematic of the system is Meshes in excess of 20,000 nodes are utilized for
shown in Fig. 1. the largest dimensions. Typical fluid node spacing is
The steady-state 2D continuity, momentum, en- 50 µm in the axial direction and 6 µm in the trans-
ergy, and species equations in the fluid phase and the verse direction. Typical wall node spacing is 50 µm in
steady-state 2D energy equation in the solid phase are the axial direction and 20 µm in the transverse direc-
discretized using a finite-volume method. Fluent 6.0 tion, where the temperature does not vary (only a few
is used to perform these calculations [22]. The flow nodes are placed in the transverse direction within the
is laminar. Previous studies of methane/air mixtures wall). However, as the mesh is nonuniform, these are
utilizing surface and gas-phase radiation have shown just representative values.
that the ignition distances for combustion reactions in The fluid viscosity, specific heat, and thermal con-
microburners were only slightly increased by radia- ductivity are calculated by a mass-fraction-weighted
tion [13]. The aspect ratio of the system is so high that average of species properties. The species specific
any surface-to-surface radiation is most likely emit- heat is calculated using a piecewise polynomial fit of
ted and absorbed at nearly the same axial location. temperature [22]. The fluid density is calculated using
Therefore, radiation is omitted from the simulations the ideal gas law.
performed in this work to focus on the effect of diffu- It has been shown that for typical materials of
sive and convective heat transport on flame stability. construction, radical quenching is important in de-
The boundary conditions used in this model are termining flame stability in microburners. In partic-
as follows. At the inlet a fixed flat velocity profile ular, simulations that were performed with complex
is assumed. For the species and energy equations, gaseous chemistry have shown that radical sticking
Danckwerts boundary conditions are employed; i.e., coefficients that are larger than 0.001 severely de-
the convective portions of the equations are fixed, and crease the flame stability of microburners [11]. Recent
the diffusive portions are calculated implicitly. At the experimental work has produced a nearly “quench-
interface between the fluid and the solid, no slip and less” wall material that is resistant to radical quench-
no normal species diffusive flux boundary conditions ing [15,16]. Thus, our focus here is on understanding
are applied. The heat flux at this interface is calcu- the overall heat-transfer characteristics in microburn-
lated using Fourier’s law and continuity in tempera- ers and developing guidelines for appropriate thermal
ture and heat flux is ensured. A symmetry boundary management that can result in more robust flame be-
condition is applied at the centerline between the two havior. One-step chemistries, determined from flame
plates. At the exit, the pressure is specified and the speeds, are a useful tool for describing flame dynam-
remaining variables are calculated assuming far-field ics and flame responses to external perturbations [23].
conditions, i.e., zero diffusive flux of species or en- In this work a reduced one-step propane combustion
ergy normal to the exit. In the bulk of the wall the chemistry is used that assumes the irreversible com-
2D energy equation is solved. The exterior/top sur- bustion of propane:
face of the wall is assumed to obey Newton’s law of
cooling, q  = h(Tw − Ta ), where q  is the heat flux, C3 H8 + 5O2 → 3CO2 + 4H2 O. (1)
h is the exterior convective heat-transfer coefficient,
Tw is the temperature at the exterior surface (an un- Consequently five species, namely C3 H8 , O2 , N2 ,
known of the problem), and Ta is the ambient temper- CO2 , and H2 O, are modeled with corresponding con-
ature, which is assumed to be 300 K. It is important servation equations. Specifically, the mechanism used
to note that all the 2D internal heat transfer within is the one proposed by Westbrook and Dryer [24],
100 D.G. Norton, D.G. Vlachos / Combustion and Flame 138 (2004) 97–107

 
rC3 H8 kgmol/(m3 s)
 
1.256 × 108 J/kgmol
= 4.836 × 109 · exp −
RT
· [CC3 H8 ]0.1 · [CO2 ]1.65 , (2)

where the concentrations are in units of kgmol/m3 .


We should remark that the one-step approximation
fails to describe several aspects, such as the possibility
for formation of partial oxidation products CO and H2
and the resulting superadiabatic flame temperature, or
the ignition temperature when ignition is kinetically
controlled [25], etc. As a result of this approxima-
tion, the flame location may be somewhat inaccurate.
Therefore, the results presented here should be used
as a guide to understand trends rather than an accu-
rate prediction. Fig. 2. Contours of (a) temperature [K], (b) reaction rate
The conservation equations were solved implic- [kgmol/(m3 s)], and (c) conversion for L = 600 µm, Lw =
itly with a 2D steady-state segregated solver using an 200 µm, kw = 3 (W/m)/K, h = 10 (W/m2 )/K, Vinlet =
under-relaxation method. The segregated solver first 0.5 m/s, and a stoichiometric feed (not to scale, and reflec-
solves the momentum equation, then the continuity tion of symmetry is used for easier visualization).
equation, and then updates the pressure and mass flow
rate. The conservation equations are then checked is consistent with the experimental observations by
for convergence. Convergence is determined from the Churchill of a large-diameter TSB [17].
residuals of the conservation equations as well as the Despite the small scale, there are significant tem-
difference (the L2 norm) between subsequent itera- perature and species gradients within the fluid near
tions of the solution. The pressure was discretized the reaction zone. These gradients necessitate the use
using a “Standard” method. The pressure–velocity of an elliptic model, as axial diffusion of species and
coupling was discretized using the “Simple” method. energy cannot be neglected. Within the walls there
The momentum, species, and energy equations were exist axial temperature gradients but near-zero trans-
discretized using a first-order upwind approximation. verse temperature gradients (only a handful of nodes
Details about these schemes can be found in [22].
are used within the wall in the transverse direction).
The simulations were performed on a Beowulf
Thus, when the wall temperature is displayed in the
cluster consisting of 58 Pentium IV processors and
graphs below, only the exterior wall temperature is
58 GB of RAM. When parallel processing was used,
shown.
the message passing interface (MPI) was used to
In order to understand how to design microburners
transmit information between nodes. In order to
with enhanced stability and robustness, it is necessary
achieve convergence as well as compute extinction
to understand the extinction and blowout processes.
points, natural parameter continuation was imple-
Fig. 3 shows reaction-rate profiles on the centerline
mented. The calculation time of each simulation var-
for three different cases, a stable microflame, one near
ied between 30 min and several hours, depending on
blowout, and one near extinction. Qualitatively, as a
the difficulty of the problem and the initial guess.
microburner approaches extinction the reaction zone
shifts downstream slightly and is broadened, whereas
3. Microflame characteristics the maximum reaction rate decreases. The reaction
is quenched without leaving the microburner. On the
Fig. 2 shows contour plots of the temperature, re- other hand, when blowout occurs, the reaction zone
action rate, and propane conversion for a typical set shifts significantly downstream. In both cases, loss of
of operating parameters. The entire microburner is flame stability is caused by a lack of upstream heat
shown. The flame stabilizes in the center between the transfer to the incoming reactants. The primary dif-
two plates. The reaction starts at the wall and trav- ference between the two behaviors is that in blowout
els towards the center as the flow goes downstream. more heat leaves in the form of a hot exit gas, whereas
Combustion occurs very rapidly, consuming most of in extinction, the excess heat loss occurs through the
the propane in a very small region. Complete con- walls to the surroundings. This distinction is not al-
version is achieved, and a significant temperature rise ways sharp. We should note that once the reaction
is observed due to the exothermicity of the reaction. zone shifts approximately past half the length of the
The narrow flame front observed in the simulations burner, the far-field boundary conditions at the exit
D.G. Norton, D.G. Vlachos / Combustion and Flame 138 (2004) 97–107 101

Fig. 3. Reaction rate vs axial displacement for three typi-


cal cases, a stabilized microflame, one near blowout, and
one near extinction from thermal losses. Thermal quenching
shifts the reaction downstream slightly, broadens the reac-
tion zone a bit, and reduces the maximum reaction rate,
whereas blowout shifts the reaction zone downstream sig-
nificantly without decreasing the reaction rate.

may no longer describe the system properly. As a


result, the blowout critical conditions are less accu-
rate. To overcome the accuracy problem for a fixed
microburner length, one needs to experimentally mea-
sure the exit conditions and impose them as boundary
Fig. 4. (a) Flame location vs wall thermal conductivity for
conditions. different external heat-transfer coefficients. Low wall ther-
mal conductivities cause the flame to shift downstream. In-
creasing wall thermal conductivity has little effect on flame
4. Role of wall thermal conductivity and external location unless there are significant external heat losses.
heat loss in flame stability The parameters used are Vinlet = 0.5 m/s, L = 600 µm,
Lw = 200 µm, and a stoichiometric feed. (b) Flame loca-
The location of the flame shifts as a function of op- tion vs wall thermal conductivity for different microburner
erating parameters. In [13] we introduced the flame dimensions with Vinlet = 0.5 m/s, h = 35 (W/m2 )/K, and
location as a convenient criterion for the stability or a stoichiometric feed.
robustness of a microburner, defined as the axial po-
sition with the highest reaction rate. Fig. 4a shows the nition temperature, a flame is not stabilized within
flame location as a function of wall thermal conduc- the microburner [11]. Since the conductivity of the
tivity for different external heat-transfer coefficients. walls is orders of magnitude higher than that of the
As the wall thermal conductivity decreases to low fluid, heat conduction through the walls is the pri-
values, the flame location shifts downstream for all mary mechanism of upstream heat transfer. When this
external heat-transfer coefficients. For high wall ther- upstream heat transfer is limited by low wall ther-
mal conductivity and low external-heat-loss coeffi- mal conductivity, it takes a greater distance to achieve
cients, increasing wall thermal conductivity to high the preheating, resulting in the reaction zone shifting
values has a minor effect on the flame location. On downstream (left part of the curves in Fig. 4a). This
the other hand, for high external-heat-loss coefficients makes the flame less stable. For a given wall thermal
in systems with 600-µm gaps, increasing wall ther- conductivity, increasing the external heat loss coeffi-
mal conductivity shifts the reaction downstream. This cient shifts the reaction zone downstream as more of
nonlinear behavior is caused by the interaction be- the heat generated is lost to the surroundings.
tween two competing modes of heat transfer, namely The wall thermal conductivity alone does not de-
upstream heat transfer through the walls to preheat termine the relative upstream heat transfer in the sys-
the feed, and transverse heat transfer resulting in heat tem. The wall thickness and the gap distance also
loss to the surroundings. The former is critical for play an important role. Fig. 4b shows the effect of
ignition and flame stabilization in microchannels, as conductivity on the flame location for different gap
it allows preheating of the feed without the need for distances and wall thickness. The flame location for a
an external preheater. If the upstream heat transfer is 1200-µm gap distance, which borders on mesoscale,
insufficient to increase the fluid temperature to the ig- shows behavior qualitatively different from that of
102 D.G. Norton, D.G. Vlachos / Combustion and Flame 138 (2004) 97–107

the 600-µm gap distance. As the gap distance in-


creases, the time scales for heat transfer from the
reaction zone to the walls and from the hot walls to
the inlet reactants increases because of the increased
length scale. As a result of the latter, the flame lo-
cation occurs further downstream and more conduc-
tive materials are needed for stable operation. As a
result of the former, the system is more robust to ex-
terior heat losses. In particular, for highly conductive
materials and large external heat-transfer coefficients
(e.g., 35 (W/m2 )/K), the flame location does not shift
downstream with increasing wall thermal conductiv-
ity (see flat region for the 1200-µm case); i.e., the
larger gap makes the burner very robust with respect
Fig. 5. Wall outer edge temperature profiles for different wall
to heat losses. thermal conductivities. Low wall thermal conductivities re-
Aside from the gap distance, the wall thickness sult in large axial wall-temperature gradients and high max-
is another key factor in microburner design. Fig. 4b imum temperatures. High wall thermal conductivity leads
shows that with 400-µm-thick walls, the minimum to uniform temperature profiles without hotspots. The pa-
wall thermal conductivity allowable is approximately rameters are L = 600 µm, Lw = 200 µm, Vinlet = 0.5 m/s,
half of the minimum wall thermal conductivity allow- a stoichiometric feed, and h = 10 (W/m2 )/K.
able with 200-µm-thick walls, while the burner is very
robust even for highly conductive materials. When not be inert to radical quenching. A more reasonable
the wall thickness is doubled from 200 to 400 µm, solution would be thicker walls of a more inert mate-
the amount of heat transferred upstream for a given rial that may have a lower thermal conductivity.
wall thermal conductivity is roughly doubled. Over- Parametric continuation is used to move from
all, thicker walls add mechanical and thermal stability one stationary solution to another. When the solu-
(at the expense, of course, of increased weight). tion reaches a turning point or blowout occurs, this is
The flame locations with respect to the entrance denoted as a critical point. Knowledge of critical para-
calculated in these simulations are significantly short- meter values of the external heat transfer coefficient,
er than those observed in the TSB of Churchill, i.e., wall thermal conductivity, feed composition, and flow
O(1 mm) here vs O(150 mm) in [17]. This can be velocity gives a better understanding of the important
attributed to the substantially faster thermal feedback factors controlling flame stability. These critical val-
loop of microscale systems. The primary difference ues are useful as guides, but actual values will vary
likely stems from the change in the transverse time depending on the system (e.g., dimensions) of inter-
scales for energy diffusion. If so, the ratio of time est.
scales would be (10 mm)2 /(0.6 mm)2 ∼ 280, which Fig. 6 shows the critical external heat loss co-
is comparable to the aforementioned ratio of flame lo- efficient as a function of wall thermal conductivity.
cations. These bell-shaped envelopes separate the region of
Aside from the flame location, the material ther- self-sustained combustion below the curve from the
mal conductivity affects the temperature profile within region above the curve where combustion cannot be
the wall and the possibility of hot spots. Fig. 5 shows self-sustained. The conductivity of several materials
the temperature profiles for the outer edge of the wall is also indicated by arrows. There exists a critical
for different material thermal conductivities. For low- wall thermal conductivity for propane/air mixtures,
wall-thermal-conductivity materials, significant axial at ∼ 0.1 (W/m)/K, below which combustion cannot
temperature gradients are observed. Hotspot temper- be self-sustained, even with insulating walls. When
atures in excess of 2000 K can occur, an undesir- the wall thermal conductivity increases from low val-
able situation, as it exceeds the maximum operating ues, the allowable-heat-loss coefficient first increases
temperatures of most materials of construction. Ex- quickly, and then decreases and levels off in the range
ceedingly high wall temperatures are characteristic of metals or high-thermal-conductivity ceramics such
of both micro- and macroscale thermally stabilized as SiC. The allowable-heat-loss coefficient reaches a
burners [17,26]. As the wall thermal conductivity in- maximum for insulating ceramics such as silica and
creases, the wall temperature profiles become more alumina. The behavior seen for low-conductivity ma-
uniform and the wall hot spot is eliminated. Despite terials is at first counterintuitive. Highly insulating
the apparent advantages of a higher wall thermal con- materials are poor for flame stability due to the lack
ductivity for material stability, most materials that of- of a continuous ignition source, needed to preheat the
fer high conductance are metals, and therefore would cold incoming gases.
D.G. Norton, D.G. Vlachos / Combustion and Flame 138 (2004) 97–107 103

Fig. 6. Critical external heat loss coefficient vs wall ther- Fig. 7. Sensitivity analysis of the primary differences be-
mal conductivity. Typical ceramics allow maximum external tween propane and methane microflames. The reaction-rate
heat loss coefficients. Materials with lower wall thermal con- constant has a larger effect on the solution than the heat of
ductivities limit the upstream heat transfer. Materials with reaction. The parameters are L = 600 µm, Lw = 200 µm,
higher wall thermal conductivities result in enhanced heat Vinlet = 0.5 m/s, kw = 1 (W/m)/K, h = 9 (W/m2 )/K, and
transfer to the surroundings. Propane allows self-sustained a stoichiometric feed.
combustion for higher external heat loss coefficients and
more insulating materials than methane. The rest of the pa-
placing one heat of reaction with the other, we have
rameters are the same as in Fig. 4.
also accounted for the difference in densities. The
way we accomplish this in the first numerical exper-
5. Effect of fuel on flame stability iment is by changing the heat of reaction of pseudo-
propane so that the power generated upon complete
Propane’s mechanism for the loss of stabilized combustion of pseudo-propane in the microburner
combustion is qualitatively similar to that of methane, matches that of methane. For this case, the maxi-
discussed in previous work [13]. For low wall thermal mum temperature decreases, and the reaction location
conductivities the primary mode of burner instability shifts downstream. However, these changes are small
is blowout, whereas for high wall thermal conductiv- in comparison to the difference between propane and
ities it is extinction. However, propane microflames methane microflames.
are more robust than methane microflames, as shown Next, the reaction-rate constant parameters of
in Fig. 6. Note that the methane map is a subset of the pseudo-propane are changed to those of methane. For
propane map. Lower wall thermal conductivities and this case, the maximum temperature increases, and
higher exterior heat-loss coefficients are possible. the flame location shifts significantly downstream.
In order to better understand the differences be- This solution is closer to the methane solution. The
tween propane and methane microflames, a theoret- lower apparent activation energy of propane com-
ical fuel, denoted as “pseudo-propane,” was defined. bustion (∼ 126 kJ/mol for propane compared to
This is simply a sensitivity analysis or numerical ex- ∼ 203 kJ/mol for methane [24]) causes easier ig-
periment aiming at delineating the differences be- nition and upstream flame stabilization. From this
tween fuels. Pseudo-propane has all of the properties analysis we conclude that the reaction-rate constant
of propane except for a parameter that is changed parameters have the largest effect on flame location
to a value that is identical to that for methane. The and stability between different fuels. Higher hydro-
methane/air and propane/air mixture properties, such carbons, such as octane, generally have lower ignition
as thermal conductivity, specific heats, and viscosi- temperatures than methane. It is expected that they
ties, are similar, as the primary component is nitrogen would also exhibit increased stability compared to
in both cases. Therefore, the properties of primary in- methane.
terest are the heats of reaction and the reaction-rate
constants.
Fig. 7 shows the centerline temperature profile for 6. Role of inlet velocity in flame stability and
propane, methane, and two types of pseudo-propane. fuel-lean operation limit
Note that due to the difference in molecular weights
and densities of various fuel/air mixtures, the mass- The inlet velocity plays a key role in determining
flow rates of different fuel/air mixtures are different the location of the flame in the burner [13]. Fig. 8a
when the residence time is kept constant, as happens shows the flame location as a function of inlet veloc-
in our simulations. Therefore, instead of simply re- ity for several wall thermal conductivities. For high
104 D.G. Norton, D.G. Vlachos / Combustion and Flame 138 (2004) 97–107

Fig. 9. Critical velocity vs wall thermal conductivity. The


lower curve ( –"– ) represents stability loss due to insufficient
heat generation. The upper curve ( –2– ) represents blowout.
The shaded region allows stabilized combustion. The exper-
imentally determined laminar flame speed [27] is plotted as
a dashed line. Higher wall thermal conductivity allows faster
flows. The parameters are the same as in Fig. 8.

flow velocity changes (see Fig. 8b). When the gap dis-
Fig. 8. Flame location vs inlet velocity. An optimum flow
tance is doubled from 600 to 1200 µm the blowout
rate exists for flame stability. Higher wall thermal conduc-
tivities allow higher flow rates. The laminar flame speed for
velocity decreases from ∼ 1.7 to ∼ 0.8 m/s. This
a stoichiometric mixture of propane/air initially at 25 ◦ C is decrease in stability with respect to flow is due to
denoted as Vlam . The results for different wall thermal con- the increased timescales for energy diffusion between
ductivities are shown in (a). The parameters are L = 600 µm, the gas and the walls, resulting in the relative slow-
Lw = 200 µm, h = 10 (W/m2 )/K, and a stoichiomet- ing of the upstream preheating process, which shifts
ric feed. The results for different burner dimensions are the flame location downstream. In contrast, increas-
shown in (b). The parameters are h = 10 (W/m2 )/K, kw = ing the wall thickness from 200 to 400 µm increases
7.5 (W/m)/K, and a stoichiometric feed. the blowout velocity from ∼ 1.7 to ∼ 2.5 m/s while
leaving unaffected the flame stability for slow flows.
inlet velocities the location of the flame shifts down- This increased flame stability is due to the increased
stream with increasing flow rate due to the decrease area for heat flux, doubling in our example the up-
in the convective timescale (shorter residence times). stream preheating rate. When flow velocities greater
For low inlet velocities, a sharp shift of the reaction than the unconfined flame speed are required, the gap
zone downstream occurs with decreasing flow rate. distance must be small, and the wall thermal con-
This is due to the decrease in the heat generation rate. ductivity and thickness must be sufficiently high to
The external heat loss rate does not decrease as fast provide adequate thermal feedback to preheat the in-
as the heat generation rate, resulting in a reduced up- coming reactants.
stream heat-transfer rate. As a result of competition Fig. 9 shows the critical velocity envelope vs the
between increased volumetric heat released and de- wall thermal conductivity for a fixed external heat
creased residence time with increasing flow rate, there loss coefficient. The upper curve represents the high-
is a minimum in the flame location between 0.3 and velocity limit, resulting in blowout due to decreased
0.5 m/s, depending on the wall thermal conductiv- convective timescales. The lower curve represents the
ity. This minimum is near the unconfined flame speed low-velocity limit, resulting in flame stability loss
for the same composition, experimentally determined due to reduced heat generation. Between these curves
by Dugger and others to be ∼ 0.4 m/s [27,28]. The stabilized combustion is allowed, whereas outside
minimum shifts slightly toward higher flow rates for the envelope, self-sustained combustion is impossi-
higher wall thermal conductivity since the latter al- ble. Smaller wall thermal conductivities allow stabi-
lows greater upstream heat transfer to compete with lized combustion for lower flow rates. Lower flow
the faster convective flow. rates require less upstream heating and more insu-
As shown above, when conductivity is the primary lation against exterior heat losses. At the other ex-
variable, the gap distance and wall thickness play a vi- treme, higher wall thermal conductivities result in
tal role in the stabilization of the flame also when the maximum allowable flow rates (upper curve), but the
D.G. Norton, D.G. Vlachos / Combustion and Flame 138 (2004) 97–107 105

increased heat losses prohibit low flow rates (lower orders of magnitude higher than the one calculated in
curve). This relationship is important when designing this work. Furthermore, the experimental minimum
devices. When a low-power device is desired, more equivalence ratio is approximately 0.2, compared to
insulating materials should be preferred. On the other the ∼ 0.56 found in this work. The differences be-
hand, when a high-power device is desired, more con- tween the experimental and our computational results
ductive materials should be chosen. Our focus here may lie in the enhanced preheating and insulation that
has been on thermal stability. However, other material are achieved with the Swiss roll design and the gap
properties, such as allowable operating temperatures, distance, which is considerably larger in the experi-
radical sticking, and mechanical strength, along with ments. More work is needed to fully understand these
the microburner efficiency (complete conversion is differences.
found here for stoichiometric mixtures using the one-
step chemistry) should be considered when choosing
a material for construction and designing microburner 7. Nusselt number analysis
dimensions.
The ability of a burner to operate under lean con- To better understand how these microscale sys-
ditions is beneficial, as it may reduce unwanted prod- tems relate to their better-understood macroscale
ucts such as coke, carbon monoxide, and nitric oxide. counterparts, a Nusselt number (Nu) analysis was
It also reduces the operating temperature, which in performed. The Nu or dimensionless heat transfer co-
turn can increase burner lifetime. Fig. 10 shows the efficient is calculated as
lean equivalence ratio operation limit for the burner
as a function of Reynolds (Re) number calculated hL (kf ∂Tf
dy )|wall L
Nu = = (3)
based on the inlet conditions and the gap distance kf,cm (Tw − Tf,cm )kf,cm
half-height. There are two regimes of low (< 12) and
and is evaluated at a given axial displacement. Here
high (> 12) Re. For low Re, stability is lost with de-
Tf is the fluid temperature, Tf,cm is the cup mixing
creasing Re because of the diminished heat generation
fluid temperature, Tw is the wall temperature at the
rate. On the other hand, for high Re, blowout happens
wall–fluid interface, and kf,cm is the fluid cup mixing
because of decreased convective time scales. At the
thermal conductivity.
transition between these two regimes, there appears
Fig. 11 shows Nu versus axial displacement for
to be a deep minimum in the fuel-lean operation limit
two different cases, one with a flame stabilized near
and possibly a turning point over a narrow regime
the entrance and another at a higher velocity near
of Re. Arc-length continuation is however needed to
blowout. Nu exhibits strongly nonmonotonic behavior
fully characterize this situation.
with an oscillation at the reaction zone that finger-
Experimental work by Ahn et al. using a Swiss roll
prints the heat source, namely walls upstream trans-
burner with gap width 3.5 mm showed an optimum
ferring heat to the cold incoming gases and combus-
Re at a relatively shallow minimum equivalence ra-
tion chemistry downstream of the entrance heating the
tio [20]. However, this optimum Re is ∼ 1000, two
walls. In both examples, Nu approaches ∼ 4, which is

Fig. 10. Minimum allowable equivalence ratio vs Reynolds Fig. 11. Nusselt number vs axial displacement (see text) for
number. The Reynolds number was calculated using the ve- two inlet flow velocities indicated. The Nusselt number is a
locity, density, and kinematic viscosity at the inlet and the strongly nonmonotonic function of position. The parameters
gap half-width. The parameters are L = 600 µm, Lw = are L = 600 µm, Lw = 200 µm, h = 10 (W/m2 )/K, kw =
200 µm, h = 10 (W/m2 )/K, and kw = 3 (W/m)/K. 1 (W/m)/K, and a stoichiometric feed.
106 D.G. Norton, D.G. Vlachos / Combustion and Flame 138 (2004) 97–107

between the constant temperature and constant flux narrow envelope of flow rates within which combus-
values for circular tubes. tion can be stabilized. When a low-power device is
Groppi and Tronconi used 3D parabolic energy/ being designed, more insulating materials should be
species balances to study methane catalytic combus- favored to minimize external heat losses. Conversely,
tion [29]. They found very high Nu and Sh near the a high-power device would favor more conductive
entrance that eventually reach the fully developed materials.
flow values predicted by Shah [30]. Their entrance Overall, propane/air microflames are more robust
behavior is reminiscent of ours, but the downstream than methane/air ones. They allow a wider range of
behavior differs substantially between the catalytic wall thermal conductivities as well as higher external-
combustion studied previously and the homogeneous heat-loss coefficients. This enhanced stability appears
combustion studied here. Comparison of our results to be due to propane’s lower ignition temperature,
with the solutions of the Graetz–Nusselt problem for which causes the reaction front to stabilize further
two parallel plates for the special cases of constant upstream than for methane. It is expected that larger
wall temperature and constant wall heat flux [31] hydrocarbons will behave more like propane than
show that the solutions of the Graetz–Nusselt prob- methane.
lem overestimate Nu. Note that existing correlations Finally, the available heat-transfer correlations are
do not take into account fluid chemistry and ignore inadequate when homogeneous reactions are present.
axial energy diffusion. To accurately capture heat transfer in microchemical
systems, an elliptic model, such as the one presented
in this work, is necessary.
8. Conclusions

The characteristics of premixed propane/air mi- Acknowledgments


crocombustion and stability envelopes were studied.
We have found that propane/air flames can be sta- This work was supported by the Army Research
bilized in narrow channels but very careful design Office under Contract DAAD19-01-1-0582. Any
is necessary. The wall material thermal conductivity opinions, findings, and conclusions or recommenda-
plays a competing role in flame stability. Walls trans- tions expressed are those of the authors and do not
fer heat upstream for ignition of the cold incoming necessarily reflect the views of the Army Research
gases but at the same time are responsible for heat Office.
losses. Consequently, there is an optimum wall ther-
mal conductivity in terms of flame stability, which ap-
pears to be that of common ceramics such as alumina References
and silica. Despite the small scales of these systems,
large transverse gradients in temperature and species [1] L. Sitzki, K. Borer, E. Schuster, P.D. Ronney, S. Wus-
mass fractions exist in the fluid and large axial gradi- sow (Eds.), The Third Asia–Pacific Conference on
ents in temperature may exist in the walls. Regarding Combustion, Seoul, Korea, 2001.
material lifetimes, higher wall thermal conductivities [2] P. Aghalayam, D.G. Vlachos, AIChE J. 44 (9) (1998)
reduce the wall temperature gradients and hotspots 2025–2034.
[3] J.C. Ganley, E.G. Seebauer, R.I. Masel, AIChE J. 50 (4)
and should be preferred. It was also shown that the
(2004) 829–834.
burner size plays a significant role. Thicker walls en-
[4] H. Davy, Trans. R. Soc. London 107 (1817) 45–76.
able more upstream heat propagation and faster flows [5] M. Fukuda, K. Koji, M. Sakamoto, Bull.
before blowout occurs and allow less conductive ma- JSME 24 (193) (1981) 1192–1197.
terials to be used. On the other hand, increasing the [6] B. Lewis, G. von Elbe, Combustion, Flames and Explo-
gap distance from the micro- to the mesoscale offers sions of Gases, Academic Press, Orlando, FL, 1987.
the advantage of higher stability for very conductive [7] A. Linan, F.A. Williams, Fundamental Aspects of Com-
materials but decreases the stability with respect to bustion, Oxford Univ. Press, New York, 1993.
blowout, and one can hardly use ceramics. These find- [8] M. Maekawa, Combust. Sci. Technol. 11 (1975) 141–
ings point to the advantage of microscale combustion 145.
[9] S. Ono, Y. Wakuri, Bull. JSME 20 (147) (1977) 1191–
and the need for sufficiently thick walls.
1198.
It has been shown that the inlet flow velocity plays
[10] D.G. Vlachos, L.D. Schmidt, R. Aris, AIChE J. 40 (6)
a competing role in flame stability. Low flow ve- (1994) 1005–1017.
locities result in reduced power generation. On the [11] S. Raimondeau, D. Norton, D.G. Vlachos, R.I. Masel,
other hand, high flow velocities decrease the convec- Proc. Combust. Inst. 29 (2003) 901–907.
tive timescale below that of the upstream heat transfer [12] P. Aghalayam, P.-A. Bui, D.G. Vlachos, Combust. The-
through the walls. As a result, there is only a relatively ory Modeling 2 (1998) 515–530.
D.G. Norton, D.G. Vlachos / Combustion and Flame 138 (2004) 97–107 107

[13] D.G. Norton, D.G. Vlachos, Chem. Eng. Sci. 58 (2003) [23] F.A. Williams, in: M.D. Smooke (Ed.), Overview
4871–4882. of Asymptotics for Methane Flame, Springer-Verlag,
[14] C.H. Kuo, C. Eastwood, L. Sitzki, K. Borer, P.D. Ron- Berlin, 1991.
ney (Eds.), Proceedings of the Third Joint Meeting of [24] C.K. Westbrook, F.L. Dryer, Combust. Sci. Technol. 27
the U.S. Sections of the Combustion Institute, 2003. (1981) 31–43.
[15] C. Jensen, R.I. Masel, G.V. Moore, M. Shannon, Com- [25] D.G. Vlachos, L.D. Schmidt, R. Aris, Combust.
bust. Sci. Technol. (2003), in press. Flame 95 (1993) 313–335.
[16] R.I. Masel, M. Shannon, Microcombustor Having [26] C.M. Miesse, R.I. Masel, M. Short, M.A. Shannon,
Submillimeter Critical Dimensions, US06193501; The Combust. Sci. Technol. (2003), in press.
Board of Trustees of the University of Illinois, Urbana, [27] G.L. Dugger, NACA Report 1061 (1952) 105–116.
IL, 2001. [28] R.M. Fristrom, Flame structure and Processes, Oxford
[17] S.W. Churchill, Chem. Eng. Technol. 12 (1989) 249– Univ. Press, New York, 1995.
254. [29] G. Groppi, E. Tronconi, Chem. Eng. Sci. 52 (1997)
[18] S.A. Lloyd, F.J. Weinberg, Nature 251 (1974) 47–49. 3521–3526.
[19] S.A. Lloyd, F.J. Weinberg, Nature 257 (1975) 367–370. [30] R.K. Shah, in: S. Kakaç, R.K. Shah, A.E. Bergles
[20] J. Ahn, C. Eastwood, L. Sitzki, K. Borer, P.D. Ron- (Eds.), Low Reynolds Number Flow Heat Exchang-
ney, in: Proceedings of the Third Joint Meeting of the ers (Fully developed laminar flow forced convection
U.S. Sections of the Combustion Institute, Chicago, IL, in channels), Hemisphere Publishing Company, New
2003. York, 1983.
[21] P.D. Ronney, Combust. Flame 135 (4) (2003) 421–439. [31] R.K. Shah, A.L. London, Laminar Flow Forced Con-
[22] Fluent. Fluent 6.0, Lebanon, NH, 2002. vection in Ducts, Academic Press, New York, 1978.

You might also like