Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Applied Energy 130 (2014) 350–356

Contents lists available at ScienceDirect

Applied Energy
journal homepage: www.elsevier.com/locate/apenergy

Combustion of propane with Pt and Rh catalysts in a meso-scale heat


recirculating combustor
Teresa A. Wierzbicki a,b, Ivan C. Lee b, Ashwani K. Gupta a,⇑
a
Department of Mechanical Engineering, University of Maryland, 2181 Martin Hall, Campus Dr., College Park, MD 20742, USA
b
Sensors and Electron Devices Directorate, U.S. Army Research Laboratory, 2800 Powder Mill Rd., Adelphi, MD 20783, USA

h i g h l i g h t s

 Compared extinction limits of catalytic and non-catalytic combustor operation.


 Analyzed catalytic conversion, product selectivity/yield, and activation energy.
 Results used to predict the catalytic combustor performance with liquid fuel.
 Catalysts offered enhanced stable operation of combustor than without catalyst.
 Rh is more suitable for liquid fuel due to higher fuel conversion and output energy.

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

Article history: The results obtained from the combustion behavior of propane over platinum and rhodium catalysts in a
Received 9 February 2014 meso-scale heat recirculating combustor are presented. The extinction limits, conversion, product selec-
Received in revised form 21 April 2014 tivity/yield, and activation energy using the two catalysts were compared in an effort to predict their per-
Accepted 31 May 2014
formance using a liquid fuel. The extinction limits were also compared to those of non-catalytic
Available online 17 June 2014
combustion in the same combustor. The results showed that the use of a catalyst greatly expanded the
range of stable operating conditions, in terms of both extinction limits and flow rates supported. The
Keywords:
Rh catalyst was found to exhibit a higher propane conversion rate, reaching a maximum of 90.4% at stoi-
Catalytic combustion
Heat-recirculating combustor
chiometric conditions (as compared to only 61.4% offered by the Pt catalyst under lean conditions), but
Meso-scale combustion the Pt catalyst had superior CO2 selectivity for most of the examined conditions, indicating more of the
Propane oxidation heat released being used for product formation as opposed to being lost to the environment. However,
Combustion behavior despite having a higher rate of heat loss, the combustion with the Rh catalyst produced an overall higher
amount of enthalpy than the Pt due to its superior fuel conversion. The Pt catalyst also had a significantly
smaller activation energy (13.8 kJ/mol) than the Rh catalyst (74.7 kJ/mol), except at equivalence ratios
richer than U = 1.75 (corresponding to catalyst temperatures below 500 °C), where it abruptly changed
to 211.4 kJ/mol, signifying a transition from diffusion-limited reactions to kinetically limited reactions
at this point. The results reveal that Rh would be a more suitable catalyst for use in liquid-fueled
meso-scale combustors, as fuel conversion has been found to be a limiting factor for combustion stability
in these systems, and as its higher output energy allows for greater flexibility of use.
Ó 2014 Elsevier Ltd. All rights reserved.

1. Introduction implying that even devices with very low thermal-to-electric effi-
ciencies can yield higher electrical outputs than batteries, and thus
Much interest in meso-scale combustion systems has been gar- offer significant performance improvements in portable electronic
nered of late due to the benefits they offer in terms of local power devices, i.e. longer lifetime and reduced weight [1]. Liquid hydro-
production, micro thrust, and micro propulsion generation. Hydro- carbons are ideal for use in these devices, as they are very
carbon fuels have much higher specific energies than batteries, energy-dense and can be easily stored. Several studies have dem-
onstrated the successful coupling of meso-scale combustion sys-
⇑ Corresponding author. Tel.: +1 301 405 5276. tems with thermoelectric/thermophotovoltaic devices to produce
E-mail addresses: twierz@umd.edu, teresa.wierzbicki.civ@mail.mil electric power [2–6]. However, there are some problems inherent
(T.A. Wierzbicki), ivan.c.lee2.civ@mail.mil (I.C. Lee), akgupta@umd.edu (A.K. Gupta). to meso-scale combustion systems. When the system is scaled

http://dx.doi.org/10.1016/j.apenergy.2014.05.069
0306-2619/Ó 2014 Elsevier Ltd. All rights reserved.
T.A. Wierzbicki et al. / Applied Energy 130 (2014) 350–356 351

Nomenclature

A0 pre-exponential factor R universal gas constant


C/O carbon to oxygen ratio t time
Ea activation energy T temperature
k reaction rate constant V_ k volumetric flow rate of species k
[k] concentration of species k mk stoichiometric coefficient of species k
NC,k total moles of carbon in species k U equivalence ratio
NC,out total moles of carbon in combustion products
r reaction rate

down, the heat loss to heat generation (or surface area to volume) The current study examines the impact of adding a porous foam
ratio increases, which can induce flame quenching as the system catalyst to a single-turn swiss roll combustor burning a mixture of
size is reduced. A prevalent technique to diminish this effect is to propane and air. Two different catalysts, rhodium and platinum,
utilize a heat recirculating, swiss roll-style combustor, which were investigated. The objective of this study is to analyze the role
allows heat to be transferred from the products to the reactants, of Rh and Pt catalysts in terms of extinction limits (particularly
thus permitting stable combustion in channels with a characteris- how these limits compare to those of non-catalytic combustion
tic dimension smaller than the flame quenching distance. The pre- in the same reactor), conversion, selectivity, and activation energy.
heating of reactants also broadens the range of stable operating Total oxidation of propane is used as a probe reaction in this study,
conditions by extending the lean flammability limits of the fuel since oxidation reactions are essential to maximizing stability and
used [7]. Much research has been done to investigate the effects efficiency while minimizing emissions in power generation and
of geometry, scaling, operating conditions, and combustor material thrust production applications [25]. The results will help guide in
on the extinction limits and thermal performance of meso- and determining as to which catalyst would be most suitable for imple-
micro-scale swiss roll combustors [8–14]. By varying the geometry mentation in a similar combustor using liquid fuels.
and inlet conditions (i.e. equivalence ratio and Reynolds number),
meso-scale swiss roll combustors can be tailored to suit a wide
range of both power generation and thrust applications [15,16]. 2. Experimental
Shirsat and Gupta [9] studied this concept in the context of thrust
generation for such applications as nanosatellites, power genera- 2.1. Combustor and catalyst
tion, and small robotics, and reported that the optimal geometry
for flame maintenance while minimizing heat loss and maximizing A single turn heat recirculating combustor similar to that uti-
exhaust enthalpy was a single-turn combustor design. lized by Shirsat and Gupta [9] and Wierzbicki et al. [15] was used.
Another intrinsic difficulty in meso- and micro-combustion The fuel and the oxidizer were introduced to the combustion
systems is the very short residence time within the combustion chamber partially premixed, with the fuel injected downstream
chamber. At such short residence times, complete combustion from the oxidizer to prevent any flashback into the inlet reactant
of heavier hydrocarbons, especially in blended fuels, may not tubing. A step entrance to the combustion chamber was created
always occur, resulting in significant soot formation and flame to generate a vortical flow structure and local recirculation zone
instabilities [15]. A possible solution to this problem is to utilize within the chamber, recirculating both heat and active species
a catalyst in the system. Catalysts have been shown to help from the products to the reactants. The aerodynamics of this flow
stabilize liquid fuel combustion at short residence times, even structure allowed for flame stabilization to occur in homogenous
for petroleum-based jet fuels [17–19]. The high surface area to gas-phase combustion, a concept used by Shirsat and Gupta [9].
volume ratio that is characteristic of small-scale systems is also The combustor was fabricated from a 1.5 in (3.8 cm) square
ideal for use with a catalyst, as increased surface area favors cat- monolith alumina silicate block using conventional micro machin-
alytic combustion by providing more sites upon which the fuel ing tools. The material has a thermal conductivity of 1.98 W/m K in
and oxidizer can adsorb. The addition of a catalyst has been its green state, and has a low porosity relative to other more ther-
shown in many studies to widen the range of stable operating mally insulating materials, such as zirconium phosphate, which
conditions, lower the overall reactor temperatures, and increase has a thermal conductivity of 0.8 W/m K [11]. Porosity is an impor-
the hydrocarbon oxidation rates [20–22]. Spadaccini et al. [23] tant factor to consider, due to the possibility of altering the ratio of
found that the addition of a porous platinum catalyst to their fuel to oxidizer within the combustion chamber via reactants dif-
microscale gas turbine allowed for an increase in operable mass fusing into the reactor itself. Additionally, the low thermal conduc-
flow rates through the system, leading to an 8.5-fold increase in tivity of the material ensures large temperature gradients through
power density over the maximum achieved for a non-catalytic the channel walls, enabling favorable heat transfer rates from
system of similar geometry. Catalysts also play a key role in ther- products to reactants, as well as reduced heat loss from the com-
mal management—lower overall temperatures result in smaller bustion chamber to the surroundings, thus increasing the resis-
temperature gradients throughout the system, reducing the risk tance to thermal quenching [12]. After machining, the combustor
of thermal stresses. Also, since chemical reactions theoretically was heat treated for 45 min at 1050 °C at a ramp rate of 2 °C per
only occur on catalyst surfaces, the heat source location is fixed, minute. This process increased the hardness and strength of the
unlike in homogenous combustion, where the flame location alumina silicate, thus lessening the possibility of the material
may change depending on operating conditions [8]. Due to this cracking due to thermal stresses.
enhanced stability and improved thermal management, The catalysts were fabricated using an incipient wetness
integration of catalytic micro- and meso-scale combustors with impregnation technique. Solutions of Rh(NO3)3 and Pt(NO3)4 were
such energy-harnessing devices as thermoelectrics/thermophoto- deposited onto yttria-stabilized zirconia (YSZ) felts with a porosity
voltaics and steam reformers is conjectured to be much simpler of 95% and thermal conductivity of approximately 0.09 W/m K.
than the corresponding non-catalytic burners [24]. This extremely low thermal conductivity allowed for very steep
352 T.A. Wierzbicki et al. / Applied Energy 130 (2014) 350–356

combustion, propane and air were allowed to flow though the


combustor and the system was heated until the propane autoignit-
ed on the catalyst. For the homogenous combustion experiments, a
high voltage spark generator was used to ignite the fuel/air mix-
ture via ignition electrodes located in the combustion chamber.
Fig. 2 shows photographs of the heterogeneous and homogeneous
combustors. The product gases from the combustor were then fed
into a micro-gas chromatograph and a mass spectrometer for com-
position analysis.
The extinction limits were determined by stabilizing a flame
within the combustion chamber and varying the flow rates of oxy-
gen, argon, and propane until the reaction could no longer be sup-
ported. To determine the effect of equivalence ratio on the
combustion behavior, propane flow rate was held constant, and
the flow rates of oxygen and argon were varied. For each operating
condition examined, the system was allowed to reach steady state,
as indicated by constant temperatures and product concentrations.
Fig. 1. Schematic of experimental facility. The two primary parameters observed were propane conversion
(to determine the effectiveness of the two catalysts) and CO2/CO
selectivity, to determine the completeness of combustion. A kinetic
temperature gradients between the front and back faces of the cat-
analysis was also performed to determine the activation energy of
alyst. The felt was then calcinated at 700 °C for 16 h before inser-
each catalyst.
tion into the combustor.

2.2. Experimental facility and procedure 3. Results and discussion

Fig. 1 shows a schematic diagram of the experimental facility. 3.1. Extinction limits
The experimental facility was designed to allow for efficient data
acquisition on the extinction limits, thermal performance, and fuel The measured extinction limits of the catalytic and non-cata-
conversion under both the heterogeneous (catalytic) and homoge- lytic combustors are shown in Fig. 3a and b, respectively. These
neous (non-catalytic) operating conditions. Three thermal conduc- limits were defined as the points at which combustion was no
tivity mass flow controllers were used to regulate the flow rates of longer sustainable. This occurred when the flame was either blown
oxygen, argon, and propane in the combustor. A mixture of 21% O2
and 79% Ar, simulating air, was used as an oxidizer. K-type thermo-
couples were used to monitor the inlet air temperature, mixture
pre-heat temperature, product gas temperature (at the exit of the
combustion chamber), and exhaust gas temperature. In addition,
for the catalytic combustion experiments, the temperatures at
the front and back faces of the catalysts were also monitored. Care-
ful insertion of the thermocouples ensured that the tips did not
protrude too far into the channels and disturb the flow. For the het-
erogeneous combustion experiments, the catalyst was placed
directly into the combustion chamber, and blank YSZ felt was
placed in the inlet channel to serve as a heat shield. To initiate

Fig. 3. Extinction limits for (a) catalytic combustors and (b) non-catalytic
Fig. 2. Photograph of (a) catalytic combustor and (b) non-catalytic combustor. combustor.
T.A. Wierzbicki et al. / Applied Energy 130 (2014) 350–356 353

out (at the lean limits) or extinguished (at the rich limits) in the
homogeneous experiments, and when the catalyst underwent a
steep, steady drop in temperature in the heterogeneous experi-
ments. In this study, the equivalence ratio, U, is defined as the
actual carbon to oxygen (C/O) ratio divided by the stoichiometric
C/O ratio:

ðC=OÞactual
U¼ ð1Þ
ðC=OÞstoich
The catalysts were both able to sustain very lean combustion,
down to an equivalence ratio of U = 0.593 for the Rh catalyst and
U = 0.322 for Pt. The ability of the Pt catalyst to support such sig-
nificantly leaner combustion than the Rh catalyst is likely due to
Fig. 4. Propane conversion percentage in catalytic combustors.
its lower activation energy (13.8 kJ/mol for Pt vs. 74.7 kJ/mol for
Rh). This will be further discussed in Section 3.3. No absolute rich
limits were found for either catalyst, even at equivalence ratios in exception of equivalence ratios below 0.8, the Rh catalyst delivered
excess of U = 15, which is consistent with previously reported a significantly higher conversion than Pt, especially under fuel rich
studies on catalytic microcombustion with propane and air [8]. conditions, reaching a maximum of 90.4% at U = 1.0. However,
However, for super-rich combustion (U > 3) the fuel conversion under lean conditions, the conversion was found to have a very
was extremely low and the exhaust gas temperature was low, indi- steep decline. It is conjectured that this is due to the shorter con-
cating that the combustor could not be used for any practical pur- tact times at the higher flow rates that accompany lean combus-
poses at these conditions. Fig. 3a shows the lean blowout limits for tion, which result in insufficient time for complete conversion of
the Rh and Pt catalytic combustors. While the results obtained fuel. This implies that at these conditions mass transfer limitations
revealed that there is a relationship between total flow rate and for the combustion reaction on Rh may exist, as the higher flow
blowout limit for the Rh catalyst (higher flow rates can sustain rates may prevent the reactants from diffusing from the bulk flow
lower equivalence ratios), the same did not appear to be true for to the catalyst surface. With the Pt catalyst, the conversion
the Pt catalyst; for all flow rates studied the blowout limit was increased as equivalence ratio decreased, reaching a maximum of
approximately constant. This difference in trends is likely due to 61.4% at U = 0.66, the leanest condition studied and thus the point
the activation energies of the catalysts. As less fuel is added to at which the most O2 was available to oxidize the propane.
the system (i.e. less input thermal energy), the overall temperature The selectivities of CO2 and CO were investigated to evaluate
decreases due to the lower heat release from the chemical reac- the completeness of combustion, as this is a good representative
tions. At low total flow rates, very low fuel flow rates must be used measure of catalyst performance. In addition to the experimental
in order to achieve lean combustion. The Pt catalyst was able to data, thermodynamic equilibrium product gas compositions were
sustain combustion at these low temperatures due to its relatively found using CHEMKIN software, and the equilibrium CO2 and CO
low activation energy. However, due to its higher activation
energy, the Rh required a higher catalyst temperature to propagate
the combustion reactions, and at low fuel flow rates the heat
release was insufficient to sustain these higher temperatures.
The non-catalytic combustor was found to have a much smaller
operating range, as shown in Fig. 3b. Although it could sustain
combustion down to a minimum equivalence ratio of 0.425, it
could not support combustion at total flow rates above approxi-
mately 950 SCCM (cm3/min). This is consistent with previous stud-
ies, which showed that the use of a porous foam catalyst increases
the maximum flow rate supportable by the combustor [23]. Addi-
tionally, for all flow rates examined, the non-catalytic combustor
had relatively low rich extinction limits, reaching a maximum at
an equivalence ratio of only 1.75.

3.2. Exhaust gas analysis

The fuel conversion, CO2/CO selectivity, and CO2/CO yield were


calculated using:

V_ C3 H8 ;in  V_ C3 H8 ;out
%Conv ersion ¼  100% ð2Þ
V_ C H ;in
3 8

NCout;k
Selectiv ity of species k ¼  100% ð3Þ
NCout
Yield of species k ¼ Conv ersion  Selectiv ity ð4Þ

where V_ C3 H8 represents the volumetric flow rate of propane, and


N Cout represents the total moles of carbon present in the combustion
products.
Fig. 4 shows the total propane conversion under heterogeneous
conditions using the Rh and Pt catalysts. It is clear that, with the Fig. 5. Product selectivity in catalytic combustors for (a) CO2 and (b) CO.
354 T.A. Wierzbicki et al. / Applied Energy 130 (2014) 350–356

selectivities calculated from these values. As seen in Fig. 5, the Pt


catalyst had a considerably larger CO2 selectivity (and Rh a consid-
erably larger CO selectivity) for all the conditions examined, with
the exception of U < 0.8, where there was adequate excess O2 pres-
ent to completely oxidize the fuel. As CO2 has a higher enthalpy of
formation than CO, a higher CO2 selectivity is indicative of a higher
percentage of enthalpy being directed into forming products, and
thus is not lost to the environment. The general trend for both cat-
alysts is the highest CO2 selectivities under lean conditions,
decreasing to a minimum between 1.75 < U < 2, and then increas-
ing again beyond this point. It is expected that the selectivities
would be highest at lean operating conditions, as there is excess
oxygen available to form CO2, but the reason for the increase above
U = 2 is not as obvious. It is conjectured that this may be due to the
longer contact times occurring with richer conditions that allow
more time for the reaction to overcome the lack of excess oxygen.
This requires substantiation and further examination.
The equilibrium calculations did not take into account the cat-
alytic effect on product selectivity, as evidenced by the discrepan-
cies between the two data sets. As seen in Fig. 6a, the CO2 and CO
selectivities from the Rh mimic the equilibrium trend fairly well up
until U = 2, where the experimental trend diverges from the
model, with the equilibrium calculations under-predicting the CO
(and over-predicting the CO2). This over-production of CO is fairly
typical behavior of Rh, as it has been found that, at operating tem-
peratures similar to those in this study, Rh does not catalyze the
water gas shift (WGS) reaction (Eq. (5)) very well [26].
H2 O þ CO ! H2 þ CO2 ð5Þ

Fig. 7. Product yield in catalytic combustors for (a) CO2, (b) CO, and (c) H2.

The experimental data from the Pt catalyst, on the other hand,


although exhibiting a similar trend, does not mimic the shape of
the equilibrium curve. As shown in Fig. 6b, lean of U = 1, the CO
selectivity is under-predicted and the CO2 selectivity over-pre-
dicted, and the opposite is true under fuel-rich conditions.
Despite having a significantly lower CO2 selectivity, due to the
superior conversion rates of the Rh catalyst, it had a slightly higher
CO2 yield than the Pt at all examined conditions, as seen in Fig. 7a.
This indicates that, in spite of having a higher rate of heat loss to
the environment, there is an overall higher enthalpy in the exhaust
from the Rh catalyst, which is reflected in the slightly higher tem-
peratures recorded for virtually every operating condition exam-
ined. This was further validated by calculating the enthalpy in
the exhaust for both catalysts (see Fig. 8). A higher amount of
enthalpy present in the exhaust implies greater potential for use
in both thrust production and power generation applications. The
Rh catalyst also had significantly higher CO and H2 yields than
Fig. 6. Experimental (closed symbols) and equilibrium (open symbols) CO2 and CO the Pt (Fig. 7b and c), which suggests that a meso-scale combustion
selectivities for (a) Rh catalyst and (b) Pt catalyst. system with an Rh catalyst could be further optimized to produce
T.A. Wierzbicki et al. / Applied Energy 130 (2014) 350–356 355

Fig. 8. Exhaust enthalpy in catalytic combustors.

syngas, for such applications as fuel cells. Additionally, at the con-


dition of maximum conversion (90% at U = 1), the Rh system
paired with a low-temperature thermoelectric device can achieve
a power density up to 70% higher than a lithium ion battery of
equivalent mass.
The data obtained here suggests that, in terms of conversion,
selectivity, and yield, Rh is a more suitable choice than Pt for a
meso-scale combustor. Conversion is particularly important when
utilizing liquid fuels in a combustor of this scale, as the residence
time is often not sufficient to fully combust the heavier
hydrocarbons.
Fig. 9. Arrhenius plots for (a) Rh catalyst and (b) Pt catalyst.
3.3. Kinetic analysis
erage inhibits C3H8 adsorption to the catalyst surface; however, at
The rate constant governing steady-state combustion was cal-
fuel-rich conditions, the excess C3H8 is able to prevent complete
culated using:
O(s) surface coverage [8,22]. Thus, for these low-temperature,
1 d½C3 H8  fuel-rich conditions, there is more competitive adsorption on the
r¼  ð6Þ
vC H
3 8
dt Pt surface between the C3H8 and the O2, leading to kinetically lim-
ited reactions. Furthermore, the calculated value of 211.4 kJ/mol is
r ¼ k  ½C3 H8 v C3 H8  ½O2 v O2 ð7Þ in excellent agreement with the activation energy of O(s) desorp-
tion from Pt at low temperatures (213.0 kJ/mol) provided in an
where v C3 H8 and v O2 represent the stoichiometric coefficients of established mechanism [28], further supporting that, at these con-
C3H8 and O2, respectively. As argon was a non-reacting species in ditions, the rates of the reactions occurring on the Pt surface are
these experiments, and since its concentration was very large com- controlled by the chemical kinetics.
pared to those of propane and oxygen, its contribution to the overall However, for most operating conditions studied, the activation
reaction rate was neglected. For non-stoichiometric mixtures of energy of the Pt catalyst was significantly lower than that of the
propane and air, the equivalence ratio, U, was used for v C3 H8 . The Rh, indicating greater ease of ignition (i.e. less input energy) for
results were then plotted in an Arrhenius plot (Eq. (8)), shown in the system with Pt catalyst. Based on the value of the Rh catalyst
Fig. 9, and linear regression was used to determine the activation activation energy one can conjecture that the reactions occurring
energies required for steady-state combustion. in the Rh combustor are chemical kinetics-limited. However, the
Ea steep drop in propane conversion at equivalence ratios lean of stoi-
k ¼ A0 eRT ð8Þ chiometry suggests a possible influence of mass transfer limita-
As presented in the extinction limit analysis in Section 3.1, the tions. A closer examination of Fig. 9a reveals that it is not
activation energy of the Rh catalyst was relatively high (74.7 kJ/ entirely clear from the data whether the reactions are entirely
mol) and the Pt catalyst relatively low (13.8 kJ/mol) for most of kinetics-limited, or if it is a combination of kinetic and diffusion
the conditions examined. However, as seen in Fig. 9b, the slope limitations. A future study analyzing a much wider range of oper-
of the Arrhenius plot for Pt changes drastically at low temperatures ating conditions may strive to address this question thus gaining
(T < 500 °C), which corresponds to higher equivalence ratios further insights into the complex phenomena.
(U P 1.75). For these low temperatures, the activation energy
was calculated to be 211.4 kJ/mol. This drastic change signifies a 4. Conclusions
transition from mass diffusion-limited reactions at lower equiva-
lence ratios to chemical kinetics-limited reactions at higher equiv- A meso-scale, heat recirculating combustor was utilized to
alence ratios. It has been postulated that this transition with examine the gas phase and catalytic combustion behavior of pro-
respect to equivalence ratio is due to the adsorption of O(s) onto pane. Rh and Pt catalysts were used. The results showed that the
the Pt surface, since at low temperatures O(s) desorption from addition of a catalyst to the system greatly increased the stability
the catalyst surface is slow, thus reducing the catalytic activity limits, and completely eliminated the rich extinction limit up to
[27]. Additionally, at low temperatures, the high O(s) surface cov- the regions examined. The Pt catalyst was able to support signifi-
356 T.A. Wierzbicki et al. / Applied Energy 130 (2014) 350–356

cantly leaner combustion than the Rh catalyst at all the flow rates [5] Li Y, Lien Y, Chao Y, Dunn-Rankin D. Performance of a mesoscale liquid fuel-
film combustion-driven TPV power system. Prog Photovoltaics Res Appl
examined. The two catalysts were further compared to determine
2009;17:327–36.
their effectiveness, combustion behavior, and activation energies. [6] Li Y, Li H, Dunn-Rankin D, Chao Y. Enhancing thermal, electrical efficiencies of
The results showed that, although having a considerably higher a miniature combustion-driven thermophotovoltaic system. Prog
propane conversion rate for virtually all the equivalence ratios Photovoltaics Res Appl 2009;17:502–12.
[7] Jones A, Lloyd S, Weinberg F. Combustion in heat exchangers. Proc Royal Soc
examined, the Rh catalyst had an overall poorer CO2 selectivity London A: Math Phys Eng Sci 1978;60:97–115.
than Pt, indicating a higher rate of heat loss to the environment. [8] Ahn J, Eastwood C, Sitzki L, Ronney P. Gas-phase and catalytic combustion in
However, due to its superior conversion, the CO2 yield from the heat-recirculating burners. Proc Combust Inst 2005;30(2):2463–72.
[9] Shirsat V, Gupta AK. Performance characteristics of methanol and kerosene
Rh was higher than from the Pt, indicating higher overall output fueled meso-scale heat recirculating combustors. Appl Energy
enthalpy. The calculated activation energies for each catalyst 2011;88(12):5069–82.
(74.7 kJ/mol for Rh and 13.8 kJ/mol for Pt) suggested that the com- [10] Ronney P. Analysis of non-adiabatic heat-recirculating combustors. Combust
Flame 2003;135(4):421–39.
bustion reactions on Rh were largely controlled by chemical reac- [11] Vijayan V, Gupta AK. Combustion and heat transfer at meso-scale with thermal
tion kinetics and on the Pt by diffusion limitations. However, at recuperation. Appl Energy 2010;87(8):2628–39.
high equivalence ratios (U P 1.75), there is an abrupt change in [12] Vijayan V, Gupta AK. Thermal performance of a meso-scale liquid-fuel
combustor. Appl Energy 2011;88(7):2335–43.
the reaction kinetics on the Pt catalyst, likely due to the low-tem- [13] Kim N, Aizumi A, Yokomori T, Kato S, Fujimori T, Maruta K. Development and
perature competitive adsorption between C3H8 and O2 on the Pt scale effects of small swiss-roll combustors. Proc Combust Inst
surface, which signifies a transition to kinetically limited combus- 2007;31(2):3243–50.
[14] Chen C, Ronney PD. Scale and geometry effects on heat-recirculating
tion. The results revealed that higher conversion and CO2 yield
combustors. Combust Theor Model 2013;17(5):888–905.
with the Rh catalyst make it a better choice for use of this system [15] Wierzbicki TA, Lee IC, Gupta AK. Performance of synthetic jet fuels in a meso-
with a liquid fuel. This is in spite of the higher activation energy of scale heat recirculating combustor. Appl Energy 2014;118:41–7.
the Rh; conversion is a more important parameter than activation [16] Jejurkar SY, Mishra DP. Thermal performance characteristics of a microburner
for heating and propulsion. Appl Therm Eng 2011;31:521–7.
energy to consider for these applications, as the residence time in a [17] Lee IC. Rhodium supported on thermally enhanced zeolite as catalysts for fuel
meso-scale combustor is often too short to fully combust the heavy reformation of jet fuels. Catal Today 2008;136:258–65.
hydrocarbons, and the energy required to vaporize the fuel is [18] Kyritsis DC, Roychoudhury S, McEnally CS, Pfefferle LD, Gomez A. Mesoscale
combustion: a first step towards liquid fueled batteries. Exp Thermal Fluid Sci
higher than that needed for catalytic ignition. 2004;28:763–70.
[19] Kyritsis DC, Coriton B, Faure F, Roychoudhury S, Gomez A. Optimization of a
catalytic combustor using electrosprayed liquid hydrocarbons for mesoscale
Acknowledgments power generation. Combust Flame 2004;139:77–89.
[20] Pizza G, Mantzaras J, Frouzakis CE, Tomboulides AG, Boulouchos K.
Suppression of combustion instabilities of premixed hydrogen/air flames in
Research support provided to Teresa Wierzbicki by the U.S. microchannels using heterogeneous reactions. Proc Combust Inst
Army Research Laboratory is gratefully acknowledged. Thanks 2009;32:3051–8.
[21] Kaisare NS, Deshmukh SR, Vlachos DG. Stability and performance of catalytic
are also due to Dr. Vivek Shirsat for all his help and support in var- microreactors: simulations of propane catalytic combustion on Pt. Chem Eng
ious stages of this research and in the design and development of Sci 2008;63:1098–116.
heat recirculating micro-combustors. [22] Maruta K, Takeda K, Ahn J, Borer K, Sitzki L, Ronney PD, et al. Extinction limits
of catalytic combustion in microchannels. Proc Combust Inst 2002;29:957–63.
[23] Spadaccini CM, Peck J, Waitz IA. Catalytic combustion systems for microscale
gas turbine engines. J Eng Gas Turbines Power 2007;129:49–60.
References [24] Kaisare NS, Vlachos DG. A review on microcombustion: fundamentals, devices
and applications. Prog Energy Combust Sci 2012;38:321–59.
[1] Fernandez-Pello AC. Micropower generation using combustion: issues and [25] Peela NR, Zheng W, Lee IC, Karim AM, Vlachos DG. Core–shell nanocatalyst
approaches. Proc Combust Inst 2002;29:883–99. design by combining high-throughput experiments and first-principles
[2] Federici JA, Norton DG, Brüggemann T, Voit KW, Wetzel ED, Vlachos DG. simulations. ChemCatChem 2013;5:3712–8.
Catalytic microcombustors with integrated thermoelectric elements for [26] Nogare DD, Degenstein NJ, Horn R, Canu P, Schmidt LD. Modeling spatially
portable power production. J Power Sources 2006;161:1469–78. resolved data of methane catalytic partial oxidation on Rh foam catalyst at
[3] Kania T, Schilder B, Kissel T, Stephan P, Hardt S, Dreizler A. Development of a different inlet compositions and flow rates. J Catal 2011;227:134–48.
miniaturized energy converter without moving parts. Flow Turbulent Combust [27] Deshmukh SR, Vlachos DG. A reduced mechanism for methane and one-step
2013;90:741–61. rate expressions for fuel-lean catalytic combustion of small alkanes on noble
[4] Yang WM, Chou SK, Shu C, Li ZW, Xue H. Study of catalytic combustion and its metals. Combust Flame 2007;149:366–83.
effect on microthermophotovoltaic power generators. J Appl Phys D: Appl Phys [28] Deutschmann O, Schmidt R, Behrendt F, Warnat J. Numerical modeling of
2005;38:4252–5. catalytic ignition. Proc Combust Inst 1996;26(1):1747–54.

You might also like