Chemical Engineering Science 63 (2008) 1098 – 1116 www.elsevier.com/locate/ces Stability and performance of catalytic microreactors: Simulations of propane catalytic combustion on Pt Niket S. Kaisare, Soumitra R. Deshmukh, Dionisios G. Vlachos ∗ Department of Chemical Engineering, Center for Catalytic Science and Technology (CCST), University of Delaware, Newark, DE 19716, USA Received 27 February 2007; received in revised form 23 October 2007; accepted 3 November 2007 Available online 17 November 2007 Abstract A pseudo-two-dimensional (2D) model is developed to analyze the operation of platinum-catalyzed microburners for lean propane–air combustion. Comparison with computational fluid dynamics (CFD) simulations reveals that the transverse heat and mass transfer is reasonably captured using constant values of Nusselt and Sherwood numbers in the pseudo-2D model. The model also reasonably captures the axial variations in temperatures observed experimentally in a microburner with a 300 m gap size. It is found that the transverse heat and mass transport strongly depend on the inlet flow rate and the thermal conductivity of the burner solid structure. The microburner is surface reaction limited at very low velocities and mass transfer limited at high velocities. At intermediate range of velocities (preferred range of reactor operation), mass transfer affects the microburner performance strongly at low wall conductivities, whereas transverse heat transfer affects stability under most conditions and has a greater influence at high wall conductivities. At sufficiently low flow rates, complete fuel conversion occurs and reactor size has a slight effect on operation (conversion and temperature). For fast flows, propane conversion strongly depends on residence time; for a reactor with gap size of 600 m, a residence time higher than 6 ms is required to prevent propane breakthrough. The effect of reactor size on stability depends on whether the residence time or flow rate is kept constant as the size varies. Comparisons to homogeneous burners are also presented. 䉷 2007 Elsevier Ltd. All rights reserved. Keywords: Catalytic combustion; Energy; Fuel; Simulation; Microburner extinction; Propane 1. Introduction The growing interest in hydrocarbon-based sources for decentralized power generation and as replacements of existing batteries (Fernandez-Pello, 2003; National Research Council, 2004) has spawned a significant research effort in small-scale homogeneous (Kim et al., 2007; Miesse et al., 2004) and catalytic reactors (Kolb and Hessel, 2004, Norton et al., 2004, 2006, Ouyang et al., 2005; Pattekar and Kothare, 2004; Rebrov et al., 2001; Srinivasan et al., 1997). These micro chemical systems are used either for generation of hydrogen for fuel cells (Deshmukh et al., 2004; Deshmukh and Vlachos, 2005; Ganley et al., 2004; Karim et al., 2005; Kolios et al., 2005; Tonkovich et al., 2007) or for direct conversion of thermal energy released via combustion to electrical energy using ∗ Corresponding author. Tel.: +1 302 831 2830; fax: +1 302 831 1048. E-mail address: vlachos@udel.edu (D.G. Vlachos). 0009-2509/$ - see front matter 䉷 2007 Elsevier Ltd. All rights reserved. doi:10.1016/j.ces.2007.11.014 thermoelectrics (Federici et al., 2006) or thermo-photovoltaics (Yang et al., 2004). Hydrocarbon combustion is typically needed to ensure autothermal operation of these devices. Homogeneous combustion is often the preferred mode of operation at larger scales, for example, in power plants (Kiameh, 2002) and industrial reformers for hydrogen production (Udengaard et al., 1995). However, the situation rapidly changes at smaller scales due to the high surface area to volume ratio. While homogeneous combustion becomes less stable due to thermal and radical quenching (Aghalayam et al., 1998, Norton and Vlachos, 2003, 2004, Raimondeau et al., 2003), faster mass transfer can potentially result in high effective rates of catalytic reactions (Jensen, 2001). We previously studied homogeneous combustion of stoichiometric methane–air and propane–air mixtures in microburners (Norton and Vlachos, 2003, 2004) (i.e., burners with characteristic dimension less than 1 mm). The burner solid structure had a significant influence on the stability (Leach and Cadou, N.S. Kaisare et al. / Chemical Engineering Science 63 (2008) 1098 – 1116 2005), and these burners were especially prone to thermal extinction for gap sizes of ∼ 400 m or lower (Kaisare and Vlachos, 2006). However, the range of construction materials of the reactor walls that conduct sufficient heat for homogeneous ignition and still provide enough insulation to reduce the heat losses is very narrow. Even then, these stand-alone microburners are unstable at external heat loss coefficients in the regime of free convection in air. Finally, the burner operating temperatures are locally very high—often in excess of 1500 K—due to the large activation barriers and fast rates at these temperatures, raising concerns about stability of the wall materials (Deshmukh and Vlachos, 2005). In contrast to homogeneous combustion, catalytic combustion is promising at small scales. First, the transport rates as well as the catalytic surface area per unit volume increase linearly as the device size shrinks, leading to faster effective reaction rates, more stable operation, and process intensification. Second, ignition temperatures are lower. For example, hydrogen–air mixtures self-ignite and no detectable lean-burn limit is observed in a Platinum (Pt)-catalyzed microreactor of a 250-m gap (Norton et al., 2004). Third, catalytic microburners can operate at lower device temperatures and with significant heat losses, compared to their homogeneous counterparts. For example, autothermal microburner operation with less than 250 ◦ C outer wall temperature was feasible and the microburner was stable even when coupled with a thermoelectric device (Federici et al., 2006). Despite a limited number of recent studies (Karagiannidis et al., 2007; Li and Im, 2007), the fundamentals of microcatalytic combustion are not as well understood as of their homogeneous counterparts. Yet, these systems are inherently more complex. The objective of this paper is to perform a comprehensive parametric study to understand the role of operating conditions, specifically, the wall thermal conductivity, equivalence ratio, and inlet velocity in determining microburner stability and efficiency. The importance of heat and mass transfer effects is also analyzed. A computationally efficient pseudotwo-dimensional (2D) model employing reduced reaction kinetics is developed for this purpose. Insights into optimal reactor length and gap size are also developed. The organization of this paper is as follows. The next section details the development of the reactor model and the reaction kinetics. We then compare results of our model with computational fluid dynamics (CFD) simulations, followed with an assessment of our model against experimental data. The influence of wall thermal conductivity and inlet velocity on microburner operation is investigated. Finally, the effect of reactor dimensions on device performance and stability is analyzed followed by conclusions. 2. Mathematical model 2.1. Assumptions and governing equations The reactor, illustrated in Fig. 1, consists of two parallel plates coated with Pt catalyst. The nominal reactor geometry is the same as the one employed in our previous studies on catalyst 1099 h∞(Ts - T∞) insulated insulated bw Yk0 u0, Tg0 y d/2 x l Fig. 1. Schematic of the parallel plate reactor with the dash-dotted line representing the axis of symmetry. The reactor walls are coated with Pt catalyst. Table 1 Operating conditions and model parameters used in simulations Length (cm) Plate thickness (m) Gap size (m) Equivalence ratio Inlet velocity Inlet temperature (K) Wall conductivity Nusselt number Sherwood number l bw d 1, 2, 4 200 150, 300, 600, 1200 0.6, 0.75, 0.95 Varies 300 Varies 4.0 3.8 u0 Tg0 ks Nu Sh The boldface values represent nominal reactor dimensions. homogeneous microcombustion (Kaisare and Vlachos, 2006, 2007; Norton and Vlachos, 2003, 2004). Specifically, each plate is l =1 cm long and bw =200 m thick. The plates are separated by a gap size (height) of d = 600 m. In the last part of the paper, we vary the reactor dimensions from these nominal values to study size effects. The parallel plate geometry implies that the third dimension (width) of the microburner is much larger than the gap size; wherever relevant, values (e.g., volumetric flow rate) are reported assuming a 1 cm width into the plane of the paper. Propane–air mixtures of varying concentrations are fed into the reactor at room temperature (300 K). The relevant parameters used are provided in Table 1. The following assumptions are made: the flow is laminar, the pressure drop is negligible, gases follow the ideal gas law, and radiation effects are neglected due to the large aspect ratio (l/d) of the reactor (Kaisare et al., 2005). Heat losses from the walls are described using Newton’s law of cooling (as further elaborated below, the effect of surface radiation is lumped into an effective heat transfer coefficient). The reactor is modeled using a pseudo-2D model, which involves writing conservation equations in the axial direction and a lumped parameter description of transverse heat and mass transfer (in the literature, this is often referred to as 1D heterogeneous model). The governing model is as follows (the symbols are described in Notation section): j j(u) + = 0, jt jx (1) jYk j2 Yk Mk u jYk = Dk 2 + kj rgas,j + jx jt jx j − âg kmt,k (Yk − Yks ), (2) 1100 N.S. Kaisare et al. / Chemical Engineering Science 63 (2008) 1098 – 1116 0 = Mk kj rcat,j + kmt,k (Yk − Yks ), (3) j The pertinent boundary conditions are written as at inlet, x = 0: jTg jTg + uc̄p dt jx 2 j Tg = kg 2 + Hj rgas,j − âg hg (Tg − Ts ), jx u = u0 , Yk = Yk0 , Tg = Tg0 , jTs = 0; jx c̄p (9) (4) j s cs jTs j2 Ts = ks 2 + âs Hj rcat,j − â∞ h∞ (Ts − T∞ ) dt jx j + âs hg (Tg − Ts ). (5) Our code can treat variable or constant properties. In the gas-phase material and energy balances, the axial heat and mass transfer (Laplacian) terms are approximated assuming the transport coefficients to be locally constant (in order to improve numerical robustness of the model). This is unlikely to introduce modeling errors because of the high values of the Peclet number. In the literature, these terms are typically neglected for our values of Peclet number. The surface area factor , discussed below, modifies the catalytic reaction rate in Eqs. (3) and (5). The physical properties of the reactor solid structure (density, specific heat, and wall thermal conductivity) are assumed constant. In the above equations, âg = 2 d âs = â∞ = and 1 bw (6) are the surface area per unit volume computed for the gas and solid phases, respectively. The continuity Eq. (1) is modified using the ideal gas law to obtain: − jTg M̄ jYk j(u) − + = 0. Tg jt Mk jt jx (7) k The last terms in Eqs. (2)–(5) represent transverse transport. A mass fraction-based formulation of transverse mass transport is used (i.e., mass flux is computed as kmt,k (Yk − Yks )). The heat and mass transfer coefficients are computed from the Nusselt (N u) and Sherwood (Sh) numbers according to hg = Nu kg d and kmt,k = Sh Dk . d (8) The Nu and Sh numbers are discussed in the next subsection. The thermal properties (c̄p and H ) are computed using the CHEMKIN thermodynamic database (Kee et al., 1991) and the transport properties are computed using CHEMKIN gas-phase transport libraries (Kee et al., 1990). In previous CFD work, we found no gradients across the reactor wall. In this model, hg accounts for heat transfer between the fluid and the surface. The outside heat transfer coefficient, h∞ , accounts for all possible heat losses from the outside reactor surface. Dirichlet boundary conditions are imposed at the reactor entrance, and the reactor sidewalls are assumed to be insulated. Zero-flux boundary conditions are applied at the reactor outlet. jTg jYk jTs = = 0. (10) = jx jx jx The transient mass and energy conservation Eqs. (2)–(7) are solved using the method of lines until steady state is attained. Finite difference is employed to discretize the differential equations on 200 equidistant axial nodes. The resulting differential algebraic equations (DAEs) are solved using the DASPK package (Petzold, 1983) employing a numerically computed, banded Jacobian. For simulations with different reactor lengths, we increased the number of axial nodes so that the distance between adjacent nodes remains the same as that in the nominal reactor length. Natural parameter continuation is used, where a family of solutions is obtained as a function of a parameter. Each solution is obtained starting with an initial guess from the solution at the previous parameter value. at outlet, x = l: 2.2. Reaction kinetics In spite of an extensive literature on catalytic combustion of lower alkanes on noble metals (Aryafar and Zaera, 1997; Ehrhardt et al., 1992; Hayes and Kolaczkowski, 1997; Hicks et al., 1990; Lee and Trimm, 1995; Ma et al., 1996; Otto, 1989; Yao, 1980), a kinetic model valid over a wide range of parameters was not available until recently. Literature one-step mechanisms vary significantly in the activation energies and reaction orders of the fuel and oxygen. Thermodynamically consistent microkinetic mechanisms for catalytic combustion (Mhadeshwar, 2005; Mhadeshwar and Vlachos, 2005), which have been validated over a wide range of operating conditions, are desirable but computationally demanding for reactor design and optimization. We recently developed one-step kinetic rate expressions for catalytic combustion of lean lower alkane–air mixtures on noble metals via a posteriori model reduction of detailed microkinetic models (Deshmukh and Vlachos, 2007). Adsorbed oxygen is the most abundant surface species under fuel lean burn conditions and dissociative adsorption of the alkane is the rate-determining step (RDS). Consequently, propane total oxidation on Pt: C3 H8 + 5O2 → 3CO2 + 4H2 O has the following rate expression: Cs,C3 H8 kCads 3 H8 2 . ads des 1 + kO2 Cs,O2 /kO2 rcat,C3 H8 = In the above expression, ads RT T k −E ads /RT s 0 ads kk = e k and 2Mk Tref des T k −E des /RT des kk = A 0 e k . Tref The values of rate parameters are provided in Table 2. (11) N.S. Kaisare et al. / Chemical Engineering Science 63 (2008) 1098 – 1116 1101 Table 2 Kinetic rate constants for lean propane combustion on Pt C3 H8 adsorption O2 adsorption O2 desorption A0 (s−1 ) or s0 E (kcal/mol) 0.06 0.0542 8.41 × 1012 0.154 0.766 −0.796 0,2, or 4a 0 Surface area factor () Site density () (mol/cm2 ) Tref (K) b 1.0 or 1.7a 2.5 × 10−9 300 Numbers in bold indicate values used in the majority of the paper. a The theoretical value of activation energy for propane adsorption is 0 kcal/mol; E ads = 4 kcal/mol and = 1.7 were determined following sensitivity C3 H8 analysis and experimental validation (see text for details). b E des = 52.8 − 2.3(T /300) − 32.0 ∗ . O O2 Activation energies required in computation of the reaction rate, according to Eq. (11), depend on the oxygen surface coverage. Since the surface coverages of all other species except oxygen are insignificant, we use an order-one, O(1), asymptotics approximation to eliminate them and the surface site balance reduces to ∗ + O∗ = 1. Thus, the oxygen coverage is obtained by solving the nonlinear equation: ads C des kO s,O2 /kO2 2 2 , 0 0.1 0.2 0.3 (cm) 0.4 0.5 (12) where the nonlinearity arises from the coverage dependent des . Specifically, the activation energy activation energy of kO 2 varies, in our case, in a linfor oxygen desorption EOdes 2 ear manner with the oxygen coverage, O∗ , with higher O∗ coverages decreasing the desorption energy due to lateral repulsive adsorbate–adsorbate interactions. The focus of this paper is on catalytic combustion, so that rgas = 0. At some of the higher temperatures reported herein, gas-phase chemistry could become important depending on the gap size (Karagiannidis et al., 2007; Norton et al., 2004). However, we have left out gas-phase chemistry in this work since we envision that temperatures of practical operation (in terms of materials stability and safety) should be sufficiently low to prevent the onset of gas-phase chemistry. 2.3. Nusselt and Sherwood numbers Transverse heat and mass transport is lumped in our pseudo2D model in the form of heat and mass transfer coefficients. Although Nu and Sh correlations have been developed before (Di Benedetto et al., 2006; Groppi et al., 1995; Gupta and Balakotaiah, 2001; Shah and London, 1978), these studies on conventional size devices often lack the close coupling between the wall and bulk-gas that gives rise to the non-monotonic Nu number behavior observed in our previous work (Norton and Vlachos, 2004). Hence, we use CFD simulations to obtain appropriate Nu and Sh values and subsequently verify the predictive capabilities of our pseudo-2D model. Fig. 2 illustrates CFD simulation results for the nominal reactor geometry and a typical set of operating conditions. An elliptic 2D model for the reactor with planar symmetry depicted 0.2 0.0 0.4 0.8 0.6 0.05 0.1 1.0 0.15 0.2 0.25 (cm) 1500 30 Cup-mixing bulk Tg 25 1200 20 Wall Ts 15 900 10 Nu Temperature (K) 1 300 K Propane conversion 1+ 850 K Nusselt or Sherwood numbers O∗ = 1 − 1400 K 600 5 Sh 0 300 0.0 0.2 0.4 0.6 0.8 Axial coordinate, x (cm) 1.0 Fig. 2. Contours of (a) temperature and (b) propane conversion computed using Fluent CFD simulations and (c) axial profiles of temperature and Nusselt and Sherwood numbers for ks = 2 W/m/K, h∞ = 20 W/m2 /K, = 0.75, and u0 = 0.5 m/s. A jump in the Nu profile is associated with a change in the role of the reactor walls from a net heat source to a net heat sink at a location indicated by the thin vertical line. The propane activation energy and ads = 0 kcal/mol and = 1.0, respectively. catalyst surface area factor are EC H 3 8 N.S. Kaisare et al. / Chemical Engineering Science 63 (2008) 1098 – 1116 3. Benchmarking of the pseudo-2D reactor model 3.1. Comparison to CFD model results Simulations were performed for a wide range of parameters using both our simplified model and CFD simulations. We found that our model is fairly accurate in comparison to the computationally intensive CFD model, as evidenced in Fig. 3. The symbols represent the CFD simulation results, with the bulk-gas quantities being the cup-mixing averaged values, and the lines represent our model. The conditions in Fig. 3 are close to device extinction. While this is just one representative result where the temperature predictions from CFD and pseudo-2D models show excellent agreement, we found that 0.04 Propane mass fraction in Fig. 1 is solved using Fluent 6.2 (2004). The kinetics of propane oxidation are incorporated using Fluent’s User Defined Functions (UDF) with ECads = 0 kcal/mol, = 1.0 and other 3 H8 kinetic constants from Table 2. Figs. 2a and b plot the contours of temperature and propane conversion, respectively, with the abscissas truncated to highlight the relevant portion of the reactor. The wall and cup-mixing averaged gas temperatures are plotted in Fig. 2c. A microburner typically exhibits three zones: the preheating, reaction, and post-reaction zones. While these zones are clearly demarcated in homogeneous microburners, the preheating and reaction zones often overlap in catalytic microburners, as seen in Fig. 2. The wall acts as a net heat source in the preheating/reaction zones due to axial recirculation, via wall conduction of the heat released by reaction, and as a net heat sink in the post-reaction zone due to heat losses to the surroundings. This dual heat sink–source nature manifests itself as a discontinuity in the Nu number profile in Fig. 2c. A similar observation was also made in our earlier work on homogeneous microburners (Norton and Vlachos, 2004). The transition in the wall role is also demarcated at the crossover between the wall and cup-mixing gas temperatures. The asymptotic N u∞ value approaches the constant temperature asymptote of the Graetz problem (i.e., Nu∞ = 3.8) in the preheating/reaction zones and the constant flux value (N u∞ =4.15) in the post-reaction zone. On the other hand, the Sh number displays a monotonic profile with an asymptotic value of Sh∞ = 3.8, which is equal to the constant temperature asymptote for the equivalent heat transfer problem. Based on these observations, two possibilities for Nu and Sh correlations were considered for the pseudo-2D model: (i) using a constant Nu value of 4.0 (intermediate to the two asymptotic values of 3.8 and 4.15) and a constant Sh value of 3.8, or (ii) fitting the Nu and Sh CFD profiles. As demonstrated in the next section, constant Nu and Sh values are sufficient to nearly quantitatively predict the CFD results. Using axially varying Nu and Sh fits did not provide much improvement. Hence, for the sake of simplicity, the second option was deemed unnecessary for the purpose of this paper. Our choice should not be interpreted as a claim that exact mass and heat transfer correlations are unimportant in microreactors; we merely claim that carefully chosen Nu and Sh values are sufficient for the conditions of our problem. Bulk YC3H8 0.03 0.02 Surface YC3H8 0.01 0 700 Wall Ts 600 Temperature (K) 1102 500 Gas Tg 400 Symbols: CFD simulations Lines: pseudo-2D model 300 0.0 0.2 0.4 0.6 0.8 1.0 Axial distance, x (cm) Fig. 3. Comparison of axial profiles of (a) propane mass fraction and (b) wall and bulk-gas temperatures obtained from CFD simulations (symbols) and pseudo-2D model (lines) near an extinction point, i.e., ks = 20 W/m/K, u0 =0.5 m/s, =0.75, and h∞ =135 W/m2 /K. The kinetic model parameters ads = 0 kcal/mol = 1.0. are EC H 3 8 the two models deviate somewhat at higher velocities and lower heat losses. However, the maximum deviation between the temperatures predicted by our model and CFD simulations was less than 10% in all cases investigated. Larger differences are seen in the surface mass fraction, but these are limited to the near entrance region and do not seem to substantially affect reactor stability. More importantly, the simplified model was able to accurately match the extinction trends predicted using CFD (e.g., CFD and pseudo-2D simulations predicted extinction at 145 and 147 W/m2 /K, respectively, for the conditions of Fig. 3). This finding indicated that we could reliably use the pseudo-2D model for assessing operating conditions for improved microburner stability and performance. 3.2. Model uncertainty to surface reaction model parameters The activation energy for dissociative adsorption of propane was obtained using the Unity Bond Index-Quadratic Exponential Potential (UBI-QEP) theory as ECads = 0 kcal/mol. 3 H8 The UBI-QEP theory has an uncertainty of ±5 kcal/mol (Shustorovich and Sellers, 1998). We previously reported (Deshmukh and Vlachos, 2007) that a value of ECads = 3 H8 4 kcal/mol yields a better prediction of the experimental data N.S. Kaisare et al. / Chemical Engineering Science 63 (2008) 1098 – 1116 Heat loss coefficient, h∞ (W/m2/K) 200 Efuel=0; η=1 150 100 Efuel=0; η=0.25 Efuel=2; η=1 50 Efuel=4; η=1 u0=0.5 m/s; =0.75 0 100 101 Wall conductivity, ks (W/m/K) 102 Fig. 4. Critical values of external heat loss coefficient vs. wall thermal conductivity for various activation energies of propane adsorption (in kcal/mol) and catalyst surface area factor, . The model is very sensitive to both of these parameters. of Garetto and Apesteguia (2000). In addition, there is an uncertainty regarding the amount of catalyst available for reaction, which is accounted in our model using the catalyst surface area factor, . This uncertainty arises because (i) the total catalyst surface area available for reaction may be greater than the geometric surface area since the precious metal is often dispersed in a porous catalytic insert or a monolith washcoat (either the insert or the washcoat is adhered to the wall); (ii) active sites may sinter; (iii) the catalyst may get deactivated; and (iv) there may be internal mass transfer limitations. The parameter lumps all these uncertainties.Therefore, we analyze the sensitivity of model predictions to the parameters ECads and . 3 H8 Fig. 4 shows the effects of these two parameters on the critical values of the external heat loss coefficient above which the burner loses stability. The behavior seen in this graph will be analyzed extensively below. As the activation energy increases or the catalyst surface area factor () decreases, the reactor stability drops drastically. The dashed lines, representing activation energies of 2.0 and 4.0 kcal/mol, indicate that even relatively small changes in the activation energy have a significant impact on reactor stability. Likewise, the region of stable combustion shrinks significantly when the surface area factor is decreased to = 0.25. From a practical standpoint, increasing the amount of catalyst, decreasing the particle size, and minimizing internal mass transfer in case of washcoats are important factors for improving microreactor stability. 3.3. Comparison to experimental data The next task is to estimate ECads and by comparing 3 H8 the model predictions with experimental data for propane–air 1103 combustion in a Pt-catalyzed microburner (Norton et al., 2006). The reactor consisted of two 790-m-thick stainless steel plates with catalytic inserts made of anodized alumina with Pt dispersed in them. The reactor length was 6 cm and the catalyst occupied the central 5 cm. Metal thermal spreaders made of stainless steel or copper were optionally mounted on a reactor plate to vary its thermal conductivity and improve the axial thermal uniformity. Three different reactor configurations are considered for model validation: (i) “no spreader” case, simulated assuming 790 m walls with ks = 35 W/m/K; (ii) “steel spreader” case, simulated assuming 2.5 mm walls with ks = 35 W/m/K; and (iii) “Cu spreader” case, simulated assuming 2.5 mm walls with ks = 216 W/m/K. The wall thermal conductivity for the first two cases is higher than that for stainless steel to account for the nut-bolts and the reactor housing. The conductivity values are just estimates that provide reasonable description of the experimental system. In all these cases, we set the value of the external heat loss coefficient to h∞ =20 W/m2 /K and the emissivity of the external surface for radiative heat losses to 0.7. The same convective and radiative heat losses were applied to the reactor end walls as well. In a typical simulation of the “no spreader” case at an equivalence ratio (dimensionless composition) of = 0.75, the heat losses via radiation were ∼ 70% of the total heat losses. Thus, the equivalent heat loss coefficient that accounts for both convective and radiative heat losses would be as high as h∞,net = 64.5 W/m2 /K (this is in agreement with values used in our previous work). In the remaining of the paper, the heat loss coefficient reported should be interpreted as an effective one since surface radiation is not explicitly accounted for to reduce the number of model parameters. For the two different values of activation energy (ECads =0 3 H8 or 4 kcal/mol), we varied the catalytic surface area factor, , so that for a feed rate of 12.2 m/s (approximately 2.0 SLPM), we obtain extinction for the “no spreader” case at = 0.65 in close agreement with experiment. Temperature profiles for ECads = 4.0 kcal/mol and = 1.7 are shown in Fig. 5. The 3 H8 value of = 1.7 is on a lower side, considering that the catalytic inserts, made of anodized alumina, were highly porous. This may be attributed to catalyst sintering or to the fact that heat losses through the mechanical fittings and sidewalls (third dimension, into the plane of the paper) were neglected in the model. One could potentially use an even higher value of ECads 3 H8 in conjunction with a larger (and arguably a more physically reasonable) value of . Nevertheless, Fig. 5 indicates that the values of ECads = 4.0 kcal/mol and = 1.7 are good estimates 3 H8 for reasonable description of experimental data. Based on these results, the values of ECads =4.0 kcal/mol and =1.7 are used 3 H8 in the remainder of this paper. 4. Effect of materials choice 4.1. Role of reactor wall thermal conductivity in stability Figs. 6a and b show the critical values of heat loss coefficient for stable catalytic combustion as a function of wall thermal conductivity. Starting with a stable steady state solution 1104 N.S. Kaisare et al. / Chemical Engineering Science 63 (2008) 1098 – 1116 120 1400 No spreader 100 Heat loss coefficient, h∞ (W/m2/K) Wall temperature, Ts (K) 1200 1000 Cu spreader 800 Steel spreader = 0.95 80 B 60 = 0.75 A C = 0.6 40 20 Catalytic region 0 600 1.0 2.0 3.0 4.0 Axial distance, x (cm) 5.0 6.0 Fig. 5. A comparison between experimental axial temperature profiles (symbols) of Norton et al. (2006) and model predictions (lines) of near-stoichiometric mixtures at an inlet flow rate of 2.0 SLPM. The reactor was 6 cm long, with a gap size of 300 m separating the 790 m steel walls and optional steel or copper thermal spreaders. for certain sets of parameters, we perform a natural parameter continuation by increasing the heat loss coefficient until a turning point is reached, beyond which the device quenches. Each such turning point is represented with a symbol and the locus (a two-parameter bifurcation diagram) of such critical points delineates the region of stable self-sustained combustion (stable below each curve and not stable above). Fig. 6a plots stability curves for different fuel equivalence ratios at u0 = 0.5 m/s, whereas Fig. 6b plots stability curves for different inlet velocities at = 0.75. A detailed discussion on the role of inlet velocity on device stability (Fig. 6b) is deferred until the next section. The device stability increases on increasing the equivalence ratio toward the stoichiometric point, as expected. The two solid lines with circular symbols in Fig. 6a are the stability curves for = 0.75: open and filled circles representing catalytic and homogeneous combustion, respectively. Clearly, Pt-catalyzed single channel microburners are significantly more stable than their homogeneous counterparts. Catalytic combustion can be sustained at large amounts of heat losses (either to the surroundings or to an endothermic device integrated with the microburner). For many operating conditions, a heat transfer coefficient in the forced convection is entirely feasible, making catalytic devices practically useful. Our previous work revealed that the walls play a dual role in determining the stability of homogeneous microburners: they are responsible for heat losses as well as the heat recirculation via wall conduction required for preheating the cold fuel–air stream to the ignition temperature (Kaisare and Vlachos, 2006, Norton and Vlachos, 2003, 2004). In 100 u0=1 m/s Heat loss coefficient, h∞ (W/m2/K) 0.0 = 0.75 Gas-phase 80 u0=2.5 m/s u0=0.5 m/s 60 40 u0=5 m/s 20 0 10-1 100 101 Wall conductivity, ks (W/m/K) 102 Fig. 6. Critical values of heat loss coefficient for stable catalytic propane–air combustion vs. wall thermal conductivity for (a) different equivalence ratios at u0 = 0.5 m/s and (b) different inlet velocities at = 0.75. The symbols denote turning points and the lines guide the eye. The filled symbols represent stability limits for homogeneous (gas-phase) combustion. The letters (A, B, and C) denote cases of three conductivities analyzed in Figs. 7 and 8. homogeneous microburners, walls made with materials of moderate thermal conductivities (such as ceramics, with ks ∼ 2.10 W/m/K) provide maximum stability (see solid circles in Fig. 6a). We have previously identified two modes of stability loss: microburner extinction at higher wall conductivities (ks 10 W/m/K) due to greater heat losses, and flame blowout at lower wall conductivities due to insufficient heat recirculation. In fact, homogeneous combustion is not sustainable even in an adiabatic microburner at very low values of wall conductivity. N.S. Kaisare et al. / Chemical Engineering Science 63 (2008) 1098 – 1116 1400 0.05 Lines: Gas Tg 1200 Temperature (K) A:ks = 0.1 W/m/K 1000 0.04 0.03 0.02 0.01 0.1 Reaction rate (mol/m2/sec) C3H8 mass fraction Symbols: Wall Ts 0 800 1105 A: 0.1 W/m/K 0.08 0.06 rcat 0.04 B: 2 W/m/K 0.02 YC3H8 Ys,C3H8 rcat 0 C: 200 W/m/K YC3H8 Ys,C3H8 600 C: ks = 200 W/m/K 0.0 B: ks = 2 W/m/K 400 rcat 0.2 0.4 0.6 0.8 Axial distance, x (cm) 1.0 YC3H8 0.0 0.2 0.4 0.6 Axial distance, x (cm) 0.8 1.0 Fig. 7. Axial profiles of wall (symbols) and gas (lines) temperatures near extinction for the three different wall conductivities, corresponding to the three cases marked in Fig. 6a. Temperature gradients near the entrance of the reactor highlight the role of external heat transfer and its effect on stability. Unlike the homogeneous case, catalytic combustion can be sustained at very low values of wall conductivity. The overall bell-like shape is still seen in Fig. 6a and the maximum appears to shift to higher wall thermal conductivity values with decreasing equivalence ratio. As the wall conductivity increases from low values, the reactor stability first increases, due to enhanced heat recirculation, and then decreases, due to higher heat dissipation and heat losses through highly conducting walls. The highest allowed heat transfer coefficient occurs for more insulating materials in catalytic microburners in comparison to homogeneous ones. While the heat loss–heat recirculation tradeoff is observed in the catalytic microburner as well, a clear demarcation between extinction and blowout does not exist. The lower effective activation energy for propane catalytic combustion results in a more spread-out reaction zone than the corresponding gas-phase counterpart. Blowout occurs because the upstream cooling of the hot reaction zone by the incoming cold feed progressively pushes the hotspot downstream until it exits the reactor. Due to the lower ignition temperature of catalytic combustion, the impact of upstream cooling is not as critical as in the homogeneous case. Additionally, as discussed below, the microburner is diffusion-limited at lower wall conductivity; the effect of upstream cooling is, therefore, even less important. 4.2. Role of reactor wall thermal conductivity in performance: hot spots and fuel conversion In order to understand the stability discussed above and the role of materials in performance, Fig. 7 compares the axial temperature profiles for the three cases, marked in Fig. 6a, which differ in wall conductivity. Fig. 8 shows the corresponding Ys,C3H8 Fig. 8. Propane mass fraction in bulk-gas (solid lines) and near the surface (symbols), and catalytic reaction rate (dashed lines) at the extinction limits for the three cases marked in Fig. 6a. The scales (depicted in the top, left corner) are the same for all three graphs. The reactor becomes more mass transfer limited at lower wall thermal conductivities. Nearly complete conversion occurs only for very low conductivity materials for these conditions. propane mass fractions and the surface reaction rates. At very low wall thermal conductivities (case A), a localized hot spot is observed and the reaction rates are much higher near this hot spot. This is obviously an undesirable situation from the thermal point of view. The wall temperature drops rapidly in the post-reaction zone due to heat losses, resulting in quenching of surface reaction. Overall, nearly complete conversion of propane is found (desirable) due to fast rates near the first half of the reactor. As the wall conductivity increases (case B), the maximum temperature at the hot spot decreases but some temperature non-uniformity is still observed. Increased axial heat recirculation through the walls moves the reaction zone upstream, resulting in higher stability than case A. Upon further increase of the wall thermal conductivity (case C), the maximum wall temperature drops further and the wall temperature becomes uniform. The temperatures in this case are ideal for coupling with energy generation devices, such as higher temperature thermoelectrics. The maximum surface reaction rate is now much lower, which results in decreased stability (compare case C to B in Fig. 6a). The critical heat loss coefficient is higher in case B than the other two cases; the last one-third of the reactor goes unutilized, in spite of incomplete propane conversion, due to low wall temperatures, as seen in Fig. 8B. In contrast, the uniform wall temperature in case C allows the entire length of the reactor to be utilized in propane conversion. At intermediate wall conductivities (e.g., ks = 20 W/m/K), the reactor behavior is qualitatively similar to the high wall conductivity case. In both cases B and C, propane conversion is not complete. Our simulations and experiments in Fig. 5 indicate that materials of high conductivity (metals and some ceramics) are N.S. Kaisare et al. / Chemical Engineering Science 63 (2008) 1098 – 1116 capable of eliminating hot spots and provide nearly isothermal wall conditions despite running very fast combustion chemistry under non-diluted conditions. This finding is an important one for robust catalyst and reactor performance and for determining kinetics from experiments. A major difference of catalytic microburners from homogeneous ones is that the much higher allowed heat loss coefficients result in much lower device temperatures that are a prerequisite for long-term stability of catalytic and housing materials. Our results regarding the effect of materials have obvious implications not only for microburners but also for monoliths and scaled-out microdevices (the latter operate more closely to adiabatic conditions than a single channel considered here, i.e., the effective heat loss coefficient is much lower in an ensemble of microsystems). 80 Low conversion Heat loss coefficient, h∞ (W/m2/K) 1106 Conversion > 95% 60 40 Tmax < 1500 K 20 4.3. Operation diagram Microburner stability is not the only one that determines the choice of reactor design and operating conditions. We have addressed this issue previously in integrated gas-phase microburners coupled with ammonia crackers producing hydrogen (Deshmukh and Vlachos, 2005). High propane conversion and low device temperature (to ensure material stability) are important performance factors. Propane conversion in excess of 95% and a maximum wall temperature of 1500 K are taken as reasonable (albeit arbitrary) thresholds in this work. Generally, both propane conversion and the maximum wall temperature decrease as the heat loss coefficient increases along a vertical operating line of Fig. 6. This is the case when the microburner is exposed to lower environmental temperatures or coupled to a strongly endothermic reaction or a thermoelectric element removing heat at an increasing rate. Conversely, as the system becomes more adiabatic, such as in a monolith or a scaled-out microchemical system, one gets the benefit of higher fuel conversion at the risk of higher temperatures. Fig. 9 presents an operation map for u0 = 1 m/s and = 0.75. The thick line represents the stability limit, with symbols denoting turning (quenching) points; the two thin lines represent the contours of 95% propane conversion and of a maximum wall temperature of 1500 K. The shaded region between the two curves is the operation region that satisfies both combustion efficiency and material stability criteria. For higher heat loss coefficients than the upper boundary, conversion drops; for lower heat loss coefficients than the lower boundary, temperatures become too high. The operation region is very narrow at low wall conductivities and expands significantly for ceramics (ks ∼ 2.10 W/m/K) and metals (ks ∼ 20 W/m/K and higher). Obviously, these boundaries can be manipulated for each material by, for example, changing residence time and/or equivalence ratio. Thus, the main lesson from this diagram is the concept that should be considered in design rather than the numbers themselves. The corresponding figures for higher velocities and equivalence ratios are skipped for brevity. As a summary, we have found that the operation region shrinks at higher velocities, due to lower conversions, as well as at higher equivalence ratios, due to material stability issues. Stability limit High temperatures 0 10-1 100 101 Wall conductivity, ks (W/m/K) 102 Fig. 9. Operation map for propane–air catalytic combustion on Pt at u0 =1 m/s and = 0.75. Thin top and bottom lines are contours for 95% propane conversion and maximum device temperature of 1500 K, respectively. The shaded region is the operation region that satisfies both propane conversion and material stability requirements. 4.4. Heat and mass transfer limitations There are several ways to characterize external heat and mass transfer limitations and their effect on microburner stability. A simple one is to determine the difference between bulk and surface quantities (driving forces). The larger the difference, the more important the corresponding transport limitation is. Our results in Fig. 8 indicate that there are substantial mass transfer limitations, especially in the reactor entrance despite the small gap size. This is in part due to the fast combustion kinetics that renders the system transport-limited. The microburner tends to be mass transfer limited at lower conductivities, where temperatures and intrinsic reaction rates are higher, and shifts toward surface reaction limited at higher wall conductivities, where temperatures and intrinsic reaction rates are lower. Transverse temperature gradients are seen in Fig. 7 for all material conductivities, especially near the entrance. This indicates that the upstream compartment connected to the microreactors needs to be considered for quantitative modeling since this will affect preheating of reactants. The relative difference between the wall and fluid temperatures in Fig. 7 suggests that heat transfer affects stability over the entire range of wall conductivities. In order to isolate the roles of transverse heat and mass transfer, Fig. 10 compares the stability of the actual microburner to two hypothetical cases of infinitely fast heat or mass transfer (simulated by increasing either Nu or Sh number, respectively, to a very large value of 105 ). Note that the vertical scale is linear, so the effect of heat transfer is as important (or more) N.S. Kaisare et al. / Chemical Engineering Science 63 (2008) 1098 – 1116 than that of mass transfer at low wall conductivities. Transverse heat transfer affects stability over the entire range of wall conductivities, but less so at intermediate wall conductivities. Increasing the heat transfer rate decreases reactor stability because greater heat transfer is tantamount to increasing the heat loss from the reactor walls to the fluid. Unlike gas-phase microburners, where the gas gets hot via combustion and heats up the wall, in catalytic combustion the heat is released at the wall and is transferred to the combustible mixture. As a result, heat transfer occurs mainly from the wall to the gas (except for small opposite gradients downstream; see Fig. 7) and heat loss not only to the surroundings but also to the fluid itself is important. In contrast to heat transfer, increasing the mass transfer rate increases reactor stability due to an increase in the effective reaction rate. The effect is small at higher wall conductivities, consistent with the gradients in mass fractions seen in Fig. 8. For more insulating walls, mass transfer becomes more important; 320 Heat loss coefficient, h∞ (W/m2/K) 240 200 Sh = ∞ 160 120 80 Actual 40 Nu = ∞ 10-1 in the limiting case of Sh=∞ and ks =0.1 W/m/K, the reactor is extremely stable because reactants diffuse “instantly” to the catalytic surface and the hot reaction zone is isolated since the heat released does not get axially dissipated. Time scales analysis, shown in Table 3, was conducted to obtain trends and better rationalize our results. The time scales for diffusion (of heat or species) between bulk-gas and wall and for reaction are given by 100 101 Wall conductivity, ks (W/m/K) 102 Fig. 10. Effect of transverse heat and mass transfer on stability. Inlet conditions: u0 = 0.5 m/s and = 0.75. d2 4D diffusion = (13) and reaction = Cg,C3 H8 rcat (T̄ , Yg,C3 H8 ).âg , (14) respectively. We use only diffusivity for these approximate estimates since its value is similar to that of thermal diffusivity. The catalytic reaction rate is computed using Eq. (11) assuming no transverse transport resistances, i.e., at a weighted average temperature: T̄ = 280 1107 (bw s cs Ts + d/2c̄p Tg ) , bw s cs + d/2c̄p and bulk-gas conditions for propane and oxygen (Yg,k ). These time scales as well as the residence time in the reactor are computed at a point located 1.5 mm from the reactor entrance. This location is sufficiently downstream to allow preheating of the reacting mixtures, but not too far downstream where the reaction has proceeded to completion. Note that the residence time reported in Table 3 is lower than the residence time computed at the reactor entrance because the velocity increases with increasing temperature and due to the volumetric increase caused by the reaction. The analysis indicates that the heat and mass diffusion time scales are within an order of magnitude to the reaction time scales, implying that close to the stability limits, there is a strong interplay of kinetics and transport. It is expected that transport limitations will be more severe away from extinction points where chemistry is typically faster. In addition, two trends are obvious: transport limitations become more important at lower wall conductivities and faster flows. The latter aspect will be discussed below. An important take-home message is that for Table 3 Residence time, time scale of transverse diffusion, and reaction time scale computed at 1.5 mm from the reactor inlet for various cases marked with letters in Figs. 6 and 11 Case Residence time diffusion reaction diffusion reaction A B C ks = 0.1 ks = 2 ks = 200 5.4 6.9 8.4 0.9 1.3 1.8 0.1 0.6 3.7 8.5 2.1 0.5 D E F u0 = 0.1 u0 = 1 u0 = 5 45 3.8 1.5 2.1 1.5 5.0 21 0.6 0.8 0.1 2.5 6.2 The time scales are in ms, the wall conductivity in W/m/K, and the inlet velocity in m/s. See main text for details. 1108 N.S. Kaisare et al. / Chemical Engineering Science 63 (2008) 1098 – 1116 fast chemistries, such as catalytic combustion, transverse transport effects in microreactors may still be significant. Residence time (s) 10 -1 5. Effect of inlet velocity = 0.75 ks = 2 80 5.2. Heat and mass transfer limitations As before, three critical points are marked in Fig. 11 for analysis: (D) extinction at low inlet velocity due to insufficient heat generation; (E) extinction at moderate inlet velocity due to large heat losses; and (F) blowout due to insufficient residence time. Fig. 12 shows temperature profiles and Fig. 13 shows species profiles along with catalytic reaction rates. The amount of chemical power input into the system depends linearly on Heat loss coefficient, h∞ (W/m2/K) 5.1. Microburner stability limits 10-3 10-4 E 60 F = 0.75 ks = 20 40 20 = 0.6 ks = 20 D 0 10-1 100 101 Inlet velocity, u0 (m/s) 102 Fig. 11. Critical values of heat loss coefficient for stable catalytic propane–air combustion as a function of inlet velocity, u0 , or inlet-based residence time for different values of wall thermal conductivity (in W/m/K) and equivalence ratio. The letters denote cases analyzed in Figs. 12 and 13. 1100 F: u0=5 m/s 900 Temperature (K) The inlet velocity affects both the residence time and the power input into a microburner via controlling the feed flow rate. As the inlet flow velocity increases, the residence time decreases but the power input increases. As a result, a maximum in stability may be expected. The flow velocity has indeed a profound effect on the stability of catalytic microburners, changing both the quantitative values and the qualitative trends of the critical heat loss coefficient vs. wall conductivity curves in Fig. 6b. The stability curves at lower velocities (e.g., u0 = 0.5 m/s) are bell-like shaped due to a tradeoff between heat recirculation and heat dissipation, as described in the previous section. At higher velocities (e.g., u0 = 2.5 m/s), the stability curves are qualitatively different though. Specifically, at higher velocities, the region of stable combustion monotonically increases with increasing wall thermal conductivity. The residence time in the reactor is no longer sufficient for complete propane combustion and the reaction zone shifts downstream with increasing flow velocity. More conducting walls recirculate heat in the reactor and preheat the incoming gases, moving the reaction zone upstream and increasing the effective reaction rate. Unlike lower velocities, there is now enough propane feed to overcome the larger heat losses encountered at higher wall conductivities. Overall, lower flow velocities are preferred for low conductivity walls whereas an optimum velocity exists for highly conductive walls. Fig. 11 depicts the critical heat loss coefficient vs. inlet velocity for different wall conductivities and equivalence ratios indicated. The inlet-based residence time is also indicated. The leaner the propane–air mixture, the less stable the microburner is. Like the homogeneous case (Kaisare and Vlachos, 2006), these curves display a typical bell-like shape: loss of microburner stability occurs due to low power input in the lowvelocity arm and to short residence time in the high-velocity arm. At a fixed equivalence ratio, insulating walls (e.g., ks = 2 W/m/K) improve stability in the low-velocity arm, due to lower heat dissipation, whereas more conducting walls (e.g., ks = 20 W/m/K) improve stability in the high velocity branch, due to higher axial heat recirculation. Further increase of the wall conductivity beyond ks = 20 W/m/K results only in a marginal change of extinction limits (data not shown). 10-2 E: u0=1 m/s 700 D: u0=0.1 m/s F: u0=5 m/s 500 Symbols: Wall T Lines: Gas T 300 0.0 0.2 0.4 0.6 Axial distance, x (cm) 0.8 1.0 Fig. 12. Axial profiles of bulk-gas (lines) and wall (symbols) temperatures for the three cases marked in Fig. 11 ( = 0.75, ks = 20 W/m/K). The wall temperature increases with inlet velocity due to a higher amount of fuel supplied per unit time. The bulk-gas temperature approaches the wall temperature further upstream at lower velocities due to higher residence time. N.S. Kaisare et al. / Chemical Engineering Science 63 (2008) 1098 – 1116 0.03 0.02 0.01 Residence time (s) rcat 10 1800 0.064 0.048 0.032 0.016 0 0 -1 10 -2 10-3 10-4 YC3H8 E: 1 m/s rcat Ys,C3H8 F: 5 m/s YC3H8 D: 0.1 m/s YC3H8 Ys,C3H8 0.0 0.2 0.4 0.6 0.8 Axial distance, x (cm) Max. wall temperature (K) 0.04 0.08 Reaction rate (mol/m2/sec) C3H8 mass fraction 0.05 1109 1500 1200 = 0.6 ks = 20 900 1.0 = 0.75; ks = 2 600 1 Ys,C3H8 rcat the inlet velocity. Hence, at low inlet velocities (case D), the device temperature is low resulting in a low reaction rate. Under these conditions, the gradients between bulk-gas and surface propane mass fractions are small and the system is surface reaction limited due to the slow chemistry (short diffusion time compared to reaction time; see Table 3) resulting from low surface temperatures. Due to high residence times, the fuel is completely converted. As the inlet velocity increases (case E), the temperature rises owing to the higher power input, and the rate of catalytic reaction increases. For faster flows, the gradients between gas and surface temperatures are substantial and the incoming cold feed starts pushing the reaction zone downstream. In case E, although the maximum in the reaction rate is downstream, it is still well within the reactor and the reaction proceeds to completion. The higher power input results in increased stability, and high heat losses are the primary cause of stability loss. At a still higher velocity of u0 = 5 m/s (case F), the residence time is lower than the diffusion time (see Table 3), resulting in poor transverse mass transport and, thus, in incomplete propane conversion. The region of maximum reaction rate is pushed toward the reactor exit; any further increase in the external heat loss coefficient results in blow out of the reaction zone and quenching of the device. We conclude our analysis by noting that the variation from a surface reaction limited regime to a mass transfer limited regime on increasing the inlet velocity (keeping all other parameters constant) is a unique feature of catalytic microcombustion (it does not have an analogue in homogeneous microburners). Interestingly, even in 600-m gap reactors (nominal case studied = 0.75; ks = 20 0.8 Propane conversion Fig. 13. Axial profiles of propane mass fractions in the bulk-gas (solid lines) and near the surface (symbols), and rates of catalytic reaction (dashed lines) for the three cases marked in Fig. 11 ( = 0.75, ks = 20 W/m/K). Close to the turning points, the reactor is kinetically limited at low inlet flow velocities (case D) due to lower temperatures and diffusion limited at higher velocities (cases E and F). The reaction zone gets blown out at high velocities (case F). = 0.75 ks = 20 = 0.6; ks = 20 0.6 0.4 = 0.75, ks = 2 0.2 Symbols: Burner extinction 0 10-1 100 101 Inlet velocity, u0 (m/s) 102 Fig. 14. (a) Maximum wall temperature and (b) propane conversion vs. inlet velocity or inlet-based residence time for different wall conductivities (W/m/K) and fuel equivalence ratios. Symbols (whenever shown) represent turning points for stable combustion. Propane breakthrough (conversion less than 95%) is observed for an inlet velocity of u0 > 2 m/s due to insufficient residence time. Shaded regions indicate an approximate range ensuring low temperatures (a) and high conversions (b). In panel a, the shaded region, to the left of the vertical dotted line, indicates the flow range within which both requirements are met. so far), transverse mass transfer limitations are significant for fast chemistries under most operating conditions. This observation has important ramifications for experimental burner design, and possible extraction of reaction kinetics from data. 5.3. Microburner efficiency and operating temperatures Fig. 14 shows the variations in the maximum wall temperature and propane conversion as a function of inlet velocity for different equivalence ratios and wall thermal conductivities, at a fixed heat loss coefficient of h∞ = 20 W/m2 /K. This corresponds to a horizontal operating line (at h∞ = 20) along Fig. 11. The symbols (whenever shown) at the ends of the curves indicate critical conditions beyond which operation is N.S. Kaisare et al. / Chemical Engineering Science 63 (2008) 1098 – 1116 6. Effect of reactor length Changing the reactor length at a constant inlet velocity changes the residence time as well as the total area available for heat loss to the surroundings. The effect of heat loss was dominant in homogeneous microburners, primarily because the reaction zone was localized, resulting in longer reactors exhibiting decreased stability (Kaisare and Vlachos, 2006). In catalytic microburners, the reaction zone can possibly span the entire length of the reactor (for example, see Fig. 13). As a result, increasing the reactor length can potentially lead to increased stability. In this section, we study the roles of increased catalyst area and heat losses, because of increasing reactor length, in microburner design. 6.1. Microburner stability The effect of reactor length on stability curves of critical heat loss coefficient vs. wall thermal conductivity is shown in Fig. 15. In Fig. 15a, the inlet velocity and thus the power input 80 70 Heat loss coefficient, h∞ (W/m2/K) impossible. Under most conditions, the temperature exhibits a bell-like shape with an expected maximum. For fast flows, the curves are close to each other, indicating that the limiting factor is the short residence time and the wall conduction time is sufficiently large. On the other hand, on the left arm of Fig. 14a (low flow velocities), the residence time is long and the conduction time scale is important: higher equivalence ratios and lower conductivities lead to higher temperatures as expected. The collapse of all three curves in Fig. 14b indicates that the conversion depends mainly on residence time with the exception of the stability limits, i.e., the minimum and maximum flow velocities delimiting stable operation. These limits depend on composition and wall material. At low inlet flow velocities (flat part of curves), as far as the flow provides enough heat to cope with heat losses, the propane conversion is complete and independent of the wall thermal conductivity, equivalence ratio, and residence time. The inlet residence time required for 95% fuel conversion is approximately 6 ms, corresponding to an inlet velocity of ∼ 1.7 m/s. This requires that we should primarily operate the catalytic combustor in the left branch of the bell-curve of Fig. 11a. Since this branch corresponds to the heat-loss-governed extinction of the microburner, insulating walls (e.g., ks = 2 W/m/K or lower) allow greater device stability than conducting walls (ks = 20 W/m/K or higher). However, the maximum wall temperatures in devices of insulating walls exceed the material stability threshold of 1500 K due to hot spot formation. The shaded regions in Figs. 14a and b are regions that satisfy the material stability limits and the combustion efficiency requirements, respectively. The range of flow for which both requirements are met is to the left of the vertical dotted line in panel a. This operation window is approximate since it depends slightly on wall conductivity and equivalence ratio. The operation window for the catalytic microburner of nominal dimensions is quite narrow. Hence, we next consider the effect of microburner dimensions on its stability. 60 l = 1 cm 50 l = 2 cm 40 30 l = 4 cm 20 10 0 80 l = 2 cm 70 Heat loss coefficient, h∞ (W/m2/K) 1110 60 50 l = 1 cm 40 l = 4 cm 30 20 10 0 10-1 100 101 Wall conductivity, ks (W/m/K) 102 Fig. 15. Effect of reactor length on the critical heat loss coefficient for =0.75 at (a) a constant inlet velocity of u0 = 0.5 m/s and (b) a constant residence time of 20 ms. As the reactor length varies, the residence time changes in panel a and the power input changes in panel b. into the system are kept constant as the length varies. Increasing the reactor length has no effect on the stability of microburners with highly insulating walls. At very low wall conductivities, the temperature in the post-reaction zone drops significantly due to heat losses and the chemistry is completed upstream (see Figs. 7 and 8 and related discussion); increasing reactor length causes no additional heat loss. As the wall conductivity increases, the axial wall temperature becomes uniform and heat losses occur through the entire reactor length. As a result, the critical heat loss coefficient decreases on increasing the reactor length for highly conducting materials, i.e., microburners made N.S. Kaisare et al. / Chemical Engineering Science 63 (2008) 1098 – 1116 6.2. Microburner efficiency and operating temperatures The maximum wall temperature and propane conversion vary with inlet velocity for a fixed heat loss coefficient and different lengths as depicted in Fig. 17. These simulations correspond to a horizontal operating line (at h∞ =20 W/m2 /K) along Fig. 16. At lower inlet velocities, nearly complete propane conversion is obtained for all the three reactor lengths. Due to the greater heat losses, the temperatures are slightly lower in the longer microburners, resulting in slight breakthrough (note the diamond in Fig. 17b). As the propane velocity increases, propane breakthrough is observed for the shortest reactor, whereas the longer reactors still exhibit complete propane conversion. On the right branch of the graph, the reactor length plays a more important role; however, conversions are incomplete. Residence time (s) -1 10-2 10 10-3 10-4 80 70 Heat loss coefficient, h∞ (W/m2/K) of highly conducting walls exhibit substantially reduced stability the longer they are. This behavior is consistent with our previous conclusion that larger heat losses through highly conducting walls are responsible for decreased microburner stability. Additional calculations for higher inlet flow velocities (e.g., 5 m/s), where breakthrough may be expected, indicate that the effect of reactor length is even less pronounced than that shown in Fig. 15a (curves are closer together; data not shown) due to a shift of the bell-like stability curves toward higher conductivities. Fig. 15b shows the effect of reactor length at a constant residence time of 20 ms. As the reactor length increases, so does the inlet velocity. Thus, the total power input into the reactor increases and offsets the higher heat losses through the longer reactor walls. The heat recirculation–heat loss tradeoff that results in a non-monotonic nature of the critical heat loss vs. wall conductivity curve is observed in all cases under these conditions. The optimal wall conductivity for maximum stability—at which the effects of greater heat loss/dissipation outweigh the effects of greater heat recirculation—shifts toward a higher value as the reactor length is increased. The higher inlet velocity in the longer reactors pushes the reaction zone downstream. This shift in the optimal wall conductivity originates because a greater amount of heat recirculation is required to preheat the cold incoming gases and keep the reaction zone upstream. Consequently, at a constant residence time, shorter reactors are more stable at lower wall conductivities, because insulating walls do not provide sufficient heat recirculation to stabilize higher velocity flows in the longer reactors. On the other hand, longer reactors are more stable at higher wall conductivities, because the higher power input results in a higher heat generation to counterbalance the increased heat losses that are responsible for device extinction. Finally, Fig. 16 shows that the shorter reactors are more stable in the heat-loss-governed extinction branch of the critical heat loss coefficient vs. inlet velocity curve. On the other hand, longer reactors provide a longer residence time, and are more stable in the right branch of the curve, wherein the residence time is lower than the diffusion time scale. Overall, the effect of reactor length on the heat loss–velocity stability graph is not dramatic. 1111 60 50 40 30 20 l = 2 cm l = 1 cm l = 4 cm 10 ks = 20 W/m/K 0 10-1 100 101 Inlet velocity, u0 (m/s) 102 Fig. 16. Effect of reactor length on critical heat loss coefficient vs. inlet velocity curves at a constant wall thermal conductivity of ks =20 W/m/K and = 0.75. The inlet-based residence time is also shown on the top horizontal axis. The above results for different reactor lengths collapse on a single curve when plotted vs. residence time, as shown in Fig. 17c. The symbols (whenever shown) represent the critical residence time for microburner quenching. The propane conversion is a strong function of the residence time. The range of residence times that provide > 95% propane conversion increases with increasing reactor length from 5–40 ms for a 1-cm-long reactor to 5–150 ms for a 4-cm-long reactor. Similar trends of maximum wall temperature, propane conversion, and reactor stability are also observed at lower values of equivalence ratio for different reactor lengths (data not shown). Given that operation should happen on the left arm of Fig. 16, our analysis suggests that there is no dramatic effect of reactor length on performance. The optimum reactor length depends on material (Fig. 15) with longer reactors exhibiting a larger range of operation with complete propane conversion. Materials stability becomes an issue at intermediate flow velocities (close to the maximum of Fig. 17a) due to higher wall temperatures; however, the equivalence ratio can be adjusted to achieve lower device temperatures. 7. Effect of gap size According to Eq. (8), the heat and mass transfer coefficients increase linearly with decreasing gap size. The former transport reduces and the latter increases microburner stability (see Fig. 10). Additionally, as the gap size decreases, the power input decreases at a constant inlet velocity. Alternatively, for a constant power input (i.e., a constant volumetric flow rate), the 1112 N.S. Kaisare et al. / Chemical Engineering Science 63 (2008) 1098 – 1116 80 1800 l = 4 cm l = 1 cm l = 2 cm 900 ks = 20 W/m/K h∞ = 20 W/m2/K 600 1 l = 2 cm 0.8 Propane conversion Heat loss coefficient, h∞ (W/m2/K) 1200 d =1200 μm 60 d = 600 μm 50 40 d = 300 μm 30 20 d = 150 μm 10 l = 1 cm 0 1 l = 4 cm 120 0.6 0.8 0.6 0.4 100 0.4 l = 1 cm l = 2 cm l = 4 cm 0.2 0.2 0 10-1 0 10-1 0 1 10 10 10 Residence time (ms) 2 100 101 Inlet velocity, u0 (m/s) 102 Fig. 17. Effect of reactor length on (a) maximum wall temperature and (b) propane conversion vs. inlet velocity. The inset (panel c) plots the propane conversion vs. residence time for the three different reactor lengths. The symbols represent critical turning points beyond which the burner quenches. The parameters are: ks = 20 W/m/K, h∞ = 20 W/m2 /K, and = 0.75. Heat loss coefficient ,h∞ (W/m2/K) Max. wall temperature (K) 70 1500 80 d = 600 μm [10 ms] 60 d = 1200 μm [20 ms] 40 d = 150 μm [2.5 ms] 20 0 10-1 inlet velocity increases and the residence time decreases with decreasing gap size. In this section, we delineate these competing effects by investigating the role of gap size. Fig. 18a shows stability curves for various gap sizes at a fixed inlet velocity. Microburners of high wall conductivities are more stable when they have wider gaps. Under these conditions where heat loss through the conducting walls dominates, the increased power input for the wider gaps, in conjunction with the reduced heat transfer from the hot walls to the cold gases dictate the effect of gap size on reactor stability. On the other hand, at lower wall conductivities, narrower microburners are more stable. In spite of the lower power input, the higher mass transfer rates substantially improve the stability of reactors of lower gap size. Fig. 18b compares the microburner stability for the alternative strategy of keeping the inlet flow rate constant (which results also in a constant Reynolds number in the parallel plate geometry). A width (third-dimension) of 1 cm is arbitrarily assumed to compute the flow rate. The narrower reactors operate d = 300 μm [5 ms] 100 101 Wall conductivity, ks (W/m/K) 102 Fig. 18. Locus of stability for different gap sizes for the same (a) inlet velocity (u0 = 0.5 m/s) and (b) inlet flow rate (360 sccm for a burner 1-cm wide) with = 0.75. The numbers in square brackets represent the respective inlet-based residence times in ms. In (a), larger gap reactors are favorable at higher wall conductivities due to both the larger power input and the reduced heat transfer from the hot walls to the cold gases. Higher mass transfer renders smaller gap size reactors more stable at low conductivities. At a fixed feed rate (panel b), narrower reactors are favorable for most practical wall materials. with lower residence times due to an increase in the inlet velocity. Analysis similar to that shown in Fig. 10 indicates that at lower residence times, the effect of mass transfer is even more profound even at high conductivities (not shown). The same conclusion was reached from the time scale analysis shown in Table 3 for the nominal sized reactor. While the time scales of heat and mass transfer between the wall and the fluid decrease with shrinking gap size, the time scale of heat loss to the surroundings remains unchanged. As a result, at very high wall N.S. Kaisare et al. / Chemical Engineering Science 63 (2008) 1098 – 1116 1800 Max. wall temperature (K) gap size, indicating that the power input is the limiting factor and diffusion is not critical. After the maximum temperature is reached, the conversion starts dropping with further increase in the flow rate. It is near the maximum temperature and on the right arm where the gap size plays an important role. Specifically, the conversion significantly increases on reducing the gap size to 300 m (or lower), as compared to the nominal reactor size. Further decreasing the gap size provides only a moderate increase in the region of stable combustion (Fig. 18b) or in the operation region (Fig. 19). Shrinking of reactor gap size comes at the cost of increased pressure drop and higher operating temperatures. Based on the results presented here, gap sizes in the range of 100.200 m are pragmatic choices for standalone catalytic microburners of propane on Pt, in contrast to the 600.1000 m range for the homogeneous (non-catalytic) counterparts (Kaisare and Vlachos, 2006). 150 μm 300 μm 1600 1400 1200 μm 600 μm 1200 1000 800 600 1 150 μm 300 μm Propane conversion 0.8 1 0.6 0.4 600 μm 0 1200 μm 600 μm 1200 μm 0.2 0.2 8. Conclusions 0.8 300 μm 0.6 0.4 150 μm 101 102 100 Residence time (ms) 0 102 103 Flow rate 1113 104 (cm3/min) Fig. 19. (a) Maximum wall temperature and (b) propane conversion vs. inlet flow rate for various reactor gap sizes with ks =20 W/m/K, h∞ =20 W/m2 K, and = 0.75 (a burner width of 1 cm is assumed to compute flow rates). The symbols (whenever shown) denote critical points for microburner quenching. Inset: propane conversion vs. inlet-based residence time for various gap sizes. conductivities, the narrower gap sizes are more stable primarily due to the increase in the mass transfer rates. This trend is intuitively expected to continue as the gap size decreases further, until the diffusion time scale becomes much lower than the intrinsic reaction time scale; any further decrease in gap size would not affect the effective reaction rate. However, Fig. 19a indicates that this may not exactly be the case since upon reduction of the gap size, the temperature keeps increasing (at sufficiently fast flows) and so does the intrinsic reaction rate constant. At very low wall conductivities, on the other hand, heat recirculation through the insulating walls is not enough to prevent blowout due to higher velocities in the narrower microburners. As a result, a crossover is observed whereby wider gaps are more stable. The effect of reactor gap size on the maximum wall temperature and conversion vs. inlet flow rate is shown in Figs. 19a and b. On the left arm of the curves, the residence time is longer than the diffusion and reaction time scales; conversion is nearly complete and the temperature keeps increasing linearly with increasing flow rate. The curves are almost independent of In this paper, we studied the stability and performance of catalytic microreactors for exothermic reactions. In the first part of the paper, a computationally efficient pseudo-2D model was developed to study catalytic combustion and was applied to propane–air mixtures in Pt-catalyzed microburners. The catalytic reactions were described by a one-step kinetic model, obtained recently by a posteriori reduction of a 104-reaction microkinetic model. The transverse heat and mass transport were described by Nusselt (Nu) and Sherwood (Sh) number correlations obtained using computational fluid dynamics (CFD) simulations. Specific results of the first part are as follows: • The catalytic microburner can be divided into a preheating, a reaction, and a post-reaction zone. The preheating and the reaction zones often overlap. • The microburner walls play a dual role: they act as a net heat source in the preheating/reaction zones, and as a net heat sink in the post-reaction zone. This is similar to homogeneous microburners with the exception that heat is generated in the gas (homogeneous) or on the wall (catalytic). • The Nu number varies as a function of axial distance in a non-monotonic manner; a sharp discontinuity in Nu number is associated with the change in the heat source–heat sink behavior of the wall. The Nu number approaches the constant temperature asymptote of Nu∞ = 3.8 in the preheating/reaction zone, and the constant heat flux asymptote of N u∞ = 4.15 in the post-reaction zone. • The Sh number profile is monotonic, with an asymptote of Sh∞ = 3.8. • With constant values of N u = 4 and Sh = 3.8, our pseudo-2D model was able to adequately capture the CFD results at a significantly lower computational cost. The resulting model was able to reasonably describe the temperature profiles observed in our prior experimental work. The thermal conductivity of the reactor solid structure, inlet velocity, and equivalence ratio have a strong influence on the 1114 N.S. Kaisare et al. / Chemical Engineering Science 63 (2008) 1098 – 1116 burner operation. Specific results of this part of our work are: • Like homogeneous burners, a tradeoff between increased heat recirculation and heat loss is also observed in catalytic microburners, resulting in an optimal wall conductivity for which the critical value of the external heat loss coefficient for microburner stability is maximum. • The optimal value of the wall thermal conductivity increases with increasing velocity. At very high velocities, the microburner stability increases monotonically with increasing wall thermal conductivity and an optimum material for stability does not exist. • Transverse mass transfer is more important at low wall conductivities, whereas heat transfer is more important at high and low wall conductivities. Increasing the mass transfer rate increases the stability, whereas increasing the heat transfer rate from the wall to the fluid decreases the stability for the entire range of wall conductivities. • The microburner tends to be diffusion limited at low wall conductivities and/or at high velocities, and surface reaction limited at low velocities. At intermediate velocities, the diffusion and intrinsic reaction rate are of the same order of magnitude. • High wall conductivity results in lower device temperature, due to better heat dissipation as well as lower conversion, especially close to extinction. • Although a lower equivalence ratio leads to a less stable burner, it provides lower device temperatures, which are important to ensure materials stability. • The microburner stability exhibits a bell-like shape with increasing inlet velocity; the left arm is controlled from low power input, it exhibits complete conversion with lower temperatures and is the most desirable operation regime. The right arm is dominated by low residence times and exhibits fuel breakthrough. Finally, we studied the effects of reactor length and gap size on catalytic microcombustion. The results depend on whether one keeps fixed the residence time or the power input as the size varies. The main points are: • For materials in the range of typical ceramics to highly conductive, longer reactors are more stable at a fixed residence time and less stable at a constant power input. Smaller gap sizes lead to enhanced stability for a fixed flow rate (power input). The opposite is true for a constant residence time. Higher pressure drops and wall temperatures should also be considered in choosing a practical (small) gap size. • For sufficiently low velocities, where conversion is complete (up to the maximum temperature), the effect of reactor length is small, and the effect of gap size is insignificant. Longer reactors and/or smaller gaps extend the flow rate region within which complete conversion is possible. In closing, we note that operation of catalytic microreactors is controlled via a large number of parameters, including materials, geometric (size), composition, the fuel itself (not studied here), and flow conditions whose effect often exhibits opposite trends resulting in optima. Both stability limits and performance characteristics are important and compound the multidimensional operation regime of these inherently complex systems. A comparison with homogeneous combustion revealed that catalytic microburners are more stable, can operate with much more insulating materials, and can operate with lower wall temperatures. While in recent years reaction engineering has mainly focused on mass transfer and kinetics, it is clear that heat transfer and heat management play a vital role at small scales. Notation â A0 bw c̄p cs C d D E h H k k ads k des kmt l M Nu rcat rgas s0 Sh t T u x X Y surface area per unit volume, m2 /m3 pre-exponential factor, s−1 wall thickness, m specific heat of gas, J/kg/K specific heat of solid, J/kg/K concentration, mol/cm3 gap size, m diffusivity, m2 /s activation energy, kcal/mol coefficient of heat transfer/loss, W/m2 /K heat of reaction, J/mol thermal conductivity, W/m/K adsorption rate constant, cm3 /mol/s desorption rate constant, s−1 mass transfer coefficient, m/s length, m molecular weight, kg/mol Nusselt number surface reaction rate, mol/m2 /s gas-phase reaction rate, mol/m3 /s sticking coefficient Sherwood number time, (s) temperature, K velocity, m/s axial coordinate, m mole fraction mass fraction Greek letters kj temperature exponent catalyst site density, mol/cm2 catalyst surface area factor surface coverage stoichiometric coefficient of species k in reaction j density, kg/m3 time scale, (s) equivalence ratio Subscripts and superscripts ads cat adsorption catalyst N.S. Kaisare et al. / Chemical Engineering Science 63 (2008) 1098 – 1116 cm des g k ref s 0 ∞ ∗ cup-mixing average desorption gas phase species index reference conditions solid wall, surface inlet conditions external/environmental conditions vacancies Acknowledgment This work was supported by the NSF (CBET-0729701). References Aghalayam, P., Bui, P.-A., Vlachos, D.G., 1998. The role of radical wall quenching in flame stability and wall heat flux: hydrogen-air mixtures. Combust Theory Modeling 2, 515–530. Aryafar, M., Zaera, F., 1997. Kinetic study of the catalytic oxidation of alkanes over nickel, palladium, and platinum foils. Catalysis Letters 48 (3–4), 173–183. Deshmukh, S.R., Vlachos, D.G., 2005. Effect of flow configuration on the operation of coupled combustor/reformer microdevices for hydrogen production. Chemical Engineering Science 60 (21), 5718–5728. Deshmukh, S.R., Vlachos, D.G., 2007. A reduced mechanism for methane and one-step rate expressions for fuel-lean catalytic combustion of small alkanes on noble metals. Combustion Flame 149, 366–383. Deshmukh, S.R., Mhadeshwar, A.B., Vlachos, D.G., 2004. Microreactor modeling for hydrogen production from ammonia decomposition on ruthenium. Industrial & Engineering Chemistry Research 43 (12), 2986–2999. Di Benedetto, A., Marra, F.S., Donsi, F., Russo, G., 2006. Transport phenomena in a catalytic monolith: effect of the superficial reaction. A.I.Ch.E. Journal 52 (3), 911–923. Ehrhardt, J.J., Colin, L., Accorsi, A., Kazmierczak, M., Zdanevitch, I., 1992. Catalytic-oxidation of methane on platinum thin-films. Sensors and Actuators B-Chemical 7 (1–3), 656–660. Federici, J.A., Norton, D.G., Bruggemann, T., Voit, K.W., Wetzel, E.D., Vlachos, D.G., 2006. Catalytic microcombustors with integrated thermoelectric elements for portable power production. Journal of Power Sources 161 (2), 1469–1478. Fernandez-Pello, A.C., 2003. Micropower generation using combustion: issues and approaches. Proceedings of the Combustion Institute 29, 883–889. Fluent, 2004. Users Manual, Version 6.2. Fluent Inc., Lebanon, NH. Ganley, J.C., Seebauer, E.G., Masel, R.I., 2004. Porous anodic alumina microreactors for production of hydrogen from ammonia. A.I.Ch.E. Journal 50 (4), 829–834. Garetto, T.F., Apesteguia, C.R., 2000. Oxidative catalytic removal of hydrocarbons over Pt/Al2 O3 catalysts. Catalysis Today 62 (2–3), 189–199. Groppi, G., Belloli, A., Tronconi, E., Forzatti, P., 1995. A comparison of lumped and distributed models of monolith catalytic combustors. Chemical Engineering Science 50 (17), 2705–2715. Gupta, N., Balakotaiah, V., 2001. Heat and mass transfer coefficients in catalytic monoliths. Chemical Engineering Science 56 (16), 4771–4786. Hayes, R.E., Kolaczkowski, S.T., 1997. Introduction to Catalytic Combustion. Gordon and Breach, London. Hicks, R.F., Qi, H.H., Young, M.L., Lee, R.G., 1990. Structure sensitivity of methane oxidation over platinum and palladium. Journal of Catalysis 122 (2), 280–294. Jensen, K.F., 2001. Microreaction engineering—is small better? Chemical Engineering Science 56 (2), 293–303. Kaisare, N.S., Vlachos, D.G., 2006. Optimal reactor dimensions for homogeneous combustion in small channels. Catalysis Today 120 (1), 96–106. 1115 Kaisare, N.S., Vlachos, D.G., 2007. Extending the region of stable homogeneous micro-combustion through forced unsteady operation. Proceedings of the Combustion Institute 31 (2), 3293–3300. Kaisare, N.S., Lee, J.H., Fedorov, A.G., 2005. Hydrogen generation in a reverse-flow microreactor: 1. Model formulation and scaling. A.I.Ch.E. Journal 51 (7), 2254–2264. Karagiannidis, S., Mantzaras, J., Jackson, G., Boulouchos, K., 2007. Hetero/homogeneous combustion and stability maps in methane-fueled catalytic microreactors. Proceedings of the Combustion Institute 31 (2), 1973–1981. Karim, A., Bravo, J., Gorm, D., Conant, T., Datye, A., 2005. Comparison of wall-coated and packed-bed reactors for steam reforming of methanol. Catalysis Today 110 (1–2), 86–91. Kee, R.J., Dixon-Lewis, G., Warnatz, J., Coltrin, M.E., Miller, J.A., 1990. A FORTRAN computer code package for the evaluation of gas phase multicomponent transport properties. Sandia National Laboratories Report, SAND86-8246. Kee, R.J., Rupley, F.M., Miller, J.A., 1991. The CHEMKIN thermodynamic data base. Sandia National Laboratories Report, SAND87-8215B. Kiameh, P., 2002. Power Generation Handbook: Selection, Applications, Operation, Maintenance. McGraw-Hill Handbooks, New York. Kim, N.I., Aizumi, S., Yokomori, T., Kato, S., Fujimori, T., Maruta, K., 2007. Development and scale effects of small Swiss-roll combustors. Proceedings of the Combustion Institute 31 (2), 3243–3250. Kolb, G., Hessel, V., 2004. Micro-structured reactors for gas phase reactions. Chemical Engineering Journal 98 (1–2), 1–38. Kolios, G., Glockler, B., Gritsch, A., Morillo, A., Eigenberger, G., 2005. Heatintegrated reactor concepts for hydrogen production by methane steam reforming. Fuel Cells 5 (1), 52–65. Leach, T.T., Cadou, C.P., 2005. The role of structural heat exchange and heat loss in the design of efficient silicon micro-combustors. Proceedings of the Combustion Institute 30, 2437–2444. Lee, J.H., Trimm, D.L., 1995. Catalytic combustion of methane. Fuel Processing Technology 42 (2–3), 339–359. Li, J., Im, H.G., 2007. Effects of dilution on the extinction characteristics of strained lean premixed flames assisted by catalytic reaction. Proceedings of the Combustion Institute 31, 1189–1195. Ma, L., Trimm, D.L., Jiang, C., 1996. The design and testing of an autothermal reactor for the conversion of light hydrocarbons to hydrogen. 1. The kinetics of the catalytic oxidation of light hydrocarbons. Applied Catalysis A-General 138 (2), 275–283. Mhadeshwar, A.B., 2005. A Hierarchical Multiscale Approach for Predictive Microkinetic Modeling of Hydrogen Production. Ph.D., Thesis, Department of Chemical Engineering, University of Delaware, Newark, DE. Mhadeshwar, A.B., Vlachos, D.G., 2005. Hierarchical multiscale mechanism development for methane partial oxidation and reforming and for thermal decomposition of oxygenates on Rh. Journal of Phyical Chemistry B 109 (35), 16819–16835. Miesse, C.M., Masel, R.I., Jensen, C.D., Shannon, M.A., Short, M., 2004. Submillimeter-scale combustion. A.I.Ch.E. Journal 50 (12), 3206–3214. National Research Council, 2004. The Hydrogen Economy: opportunities, Costs, Barriers and R&D needs. The National Academy Press, Washington, DC. Norton, D.G., Vlachos, D.G., 2003. Combustion characteristics and flame stability at the microscale: a CFD study of premixed methane/air mixtures. Chemical Engineering Science 58 (21), 4871–4882. Norton, D.G., Vlachos, D.G., 2004. A CFD study of propane/air microflame stability. Combustion Flame 138 (1–2), 97–107. Norton, D.G., Wetzel, E.D., Vlachos, D.G., 2004. Fabrication of singlechannel catalytic microburners: effect of confinement on the oxidation of hydrogen/air mixtures. Industrial & Engineering Chemistry Research 43 (16), 4833–4840. Norton, D.G., Wetzel, E.D., Vlachos, D.G., 2006. Thermal management in catalytic microreactors. Industrial & Engineering Chemical Research 45 (1), 76–84. Otto, K., 1989. Methane oxidation over Pt on gamma-alumina—kinetics and structure sensitivity. Langmuir 5 (6), 1364–1369. 1116 N.S. Kaisare et al. / Chemical Engineering Science 63 (2008) 1098 – 1116 Ouyang, X., Bednarova, L., Besser, R.S., Ho, P., 2005. Preferential oxidation (PrOx ) in a thin-film catalytic microreactor: advantages and limitations. A.I.Ch.E. Journal 51 (6), 1758–1772. Pattekar, A.V., Kothare, M.V., 2004. A microreactor for hydrogen production in micro fuel cell applications. Journal of Microelectromechanical Systems 13 (1), 7–18. Petzold, L.R., 1983. Automatic selection of methods for solving stiff and nonstiff systems of ordinary differential equations. SIAM Journal on Scientific and Statistical Computing 4, 136–148. Raimondeau, S., Norton, D.G., Vlachos, D.G., Masel, R.I., 2003. Modeling of high temperature microburners. Proceedings of the Combustion Institute 29, 901–907. Rebrov, E., de Croon, M., Schouten, J.C., 2001. Design of a microstructured reactor with integrated heat-exchanger for optimum performance of a highly exothermic reaction. Catalysis Today 69 (1–4), 183–192. Shah, R.K., London, A.L., 1978. Laminar Flow Forced Convection in Ducts. Academic Press, New York. Shustorovich, E., Sellers, H., 1998. The UBI-QEP method: a practical theoretical approach to understanding chemistry on transition metal surfaces. Surface Science Reports 31, 1–119. Srinivasan, R., Hsing, I.M., Berger, P.E., Jensen, K.F., Firebaugh, S.L., Schmidt, M.A., Harold, M.P., Lerou, J.J., Ryley, J.F., 1997. Micromachined reactors for catalytic partial oxidation reactions. A.I.Ch.E. Journal 43 (8), 3059–3069. Tonkovich, A.L.Y., Yang, B., Perry, S.T., Fitzgerald, S.P., Wang, Y., 2007. From seconds to milliseconds to microseconds through tailored microchannel reactor design of a steam methane reformer. Catalysis Today 120 (1), 21–29. Udengaard, N.R., Vannby, R., Nielsen, C.S., Kim, J.S., 1995. Steam reforming, operating experience. In: McKetta, J.J. (Ed.), Encyclopedia of Chemical Processing and Design. Marcel Dekker Inc., New York, p. 1. Yang, W.M., Chou, S.K., Shu, C., Li, Z.W., Xue, H., 2004. A prototype microthermophotovoltaic power generator. Applied Physics Letter 84 (19), 3864–3866. Yao, Y.F.Y., 1980. Oxidation of alkanes over noble-metal catalysts. Industrial & Engineering Chemistry Product Research and Development 19 (3), 293–298.
© Copyright 2025 Paperzz