Scientific paper Modeling and Finite Difference Numerical Analysis of Reaction-Diffusion Dynamics in a Microreactor Igor Plazl1* and Mitja Lakner2 1 Department of Chemical Engineering, Faculty of Chemistry and Chemical Technology, University of Ljubljana, A{ker~eva 5, SI-1000 Ljubljana, Slovenia 2 Civil and Geodetic Faculty, University of Ljubljana, Jamova 2, SI-1000 Ljubljana, Slovenia * Corresponding author: E-mail: igor.plazl@fkkt.uni-lj.si Phone: +3861 2419 512; Fax: +3861 2419 530 Received: 16-04-2009 Dedicated to the memory of the late Prof. Dr. Valentin Koloini Abstract A theoretical description with numerical experiments and analysis of the reaction-diffusion processes of homogeneous and non-homogeneous reactions in a microreactor is presented considering the velocity profile for laminar flows of miscible and immiscible fluids in a microchannel at steady-state conditions. A Mathematical model in dimensionless form, containing convection, diffusion, and reaction terms are developed to analyze and to forecast the reactor performance. To examine the performance of different types of reactors, the outlet concentrations for the plug-flow reactor (PFR), and the continuous stirred-tank reactor (CSTR) are also calculated for the case of an irreversible homogeneous reaction of two components. The comparison of efficiency between ideal conventional macroscale reactors and the microreactor is presented for a wide range of operating conditions, expressed as different Pe numbers (0.01 < Pe < 10). The numerical procedure of complex non-linear systems based on an implicit finite-difference method improved by non-equidistant differences is proposed. Keywords: Microfluidics; reaction-diffusion dynamics; microreactor; numerical analysis; non-equidistant finite differences 1. Introduction Microtechnology has uncovered new scientific solutions and challenges in a broad range of areas, from electronics, medical technology, and fuel production and processing to biotechnology, chemical industry, environmental protection, and process safety. Microscale reactors are devices whose operations depend on precisely controlled design features with characteristic dimensions from submillimeter to submicrometer. Because of the small amounts of chemicals needed and the high rate of heat and mass transfer, microscale systems are especially suited for reactions with highly toxic, flammable, and explosive reactants.1-4 In the last decade, Microreactor Technology (MRT), accepted as a new concept in chemical engineering, has impressively demonstrated the advantages of mi-crostructured devices for chemical and biochemical reactions. Numerous reactions, including many notable and in- dustrially relevant reactions, have been tried out successfully in microreactors and several hundred publications with some research review papers have appeared in peer-reviewed journals.5-10 The small length scale of microreac-tors reduces transport limitations, giving nearly gradient-less conditions desirable for the determination of reaction kinetics. The reactor miniaturization allows us to carry out reactions under more precisely controlled conditions than with conventional macroscale reactors, leading to a possibility of improved yield and selectivity of the desired products. 11-13 In the area of catalytic chemistry, microreactors are an extremely efficient tool for rapid catalyst screening and for combinatorial chemistry.14-18 In order to simulate chemical and biochemical processes in a continuous flow microreactor a coupled system of convection-diffusion-reaction in combination with hydrodynamics has been developed and many studies on microfluidic systems are available.19-26'27'28 Characterization of micro phenomena, like frictional pressure drop and viscous dissipation effects in microchannels are subject of continuous debate. The discrepancies among the work of many researchers have been summarized in review papers.29-31 The use of the Na-vier-Stokes equations appears to be appropriate for microchannel flows of liquids as long as the hydraulic diameter (Dh) of system is greater than 100 pm for conduits filled with Newtonian fluids such as water. Non-Newtonian fluid effects are expected to be important for polymeric liquids and particle suspensions flows. Wall slip effects are negligible for liquid flows in microconduits and viscous dissipation effects on the friction factor are negligible in smooth microchannels, especially for conduits with Dh > 100 pm.29 For the continuum assumption to hold in gaseous media the characteristic length should be large enough and most researchers agree on classifying the flow to be in continuum for Knudsen number (Kn) < 10-3. However, when Kn > 10, molecular level modeling is required to describe the behavior of the fluid.32,33 The present work is a theoretical and numerical outcome of an ongoing research on characterizing microreac-tor performance for different homogeneous, non-homogeneous, and heterogeneously catalyzed reactions with nonlinear reaction kinetics. The parallel flow of two immiscible fluids with the enzyme catalyzed reaction at the interphase surface in Y-shaped glass microchannel was investigated at the conditions when the continuum assumption can be applied.29 It is an attempt to analyze and to forecast the behavior of reaction-diffusion dynamics in a continuous flow microchannel with the application of a relatively simple finite difference technique. An implicit finite-difference method was used to solve the systems of 2D and 3D partial differential equations of a second order. Special attention was devoted to non-equidistant differences in order to improve the stability and accuracy of the numerical solutions. It is shown that the accuracy of the approximations can be enhanced by mesh refinement in the vicinity of the domain parts where the variations in field unknowns are expected to be significant. The numerical experiments are analyzed and examined for accuracy, stability, and theoretical consistency. To investigate the effectiveness of microreactor performance, we compared the di-mensionless outlet concentrations of a general second-order homogeneous reaction in a microchannel with that in the ideal plug-flow and continuous stirred-tank reactor for a broad range of operating conditions. 1. 1. Theoretical Background Let us consider first an irreversible homogeneous reaction taking place in the microchannel within the flow driven at a parabolic velocity profile, developed in the smallest y-dimension. Reactive components A and B enter the channel in two parallel flows as schematically presented in Figure 1. Regarding the operating conditions typical of microchannels, the laminar fluid flow in the x-axis di- rection is considered (Fig. 1). Flow in microchannel is predominantly laminar.1 Molecular effects also become more significant in microchannel when the characteristic length decreases to the point that the continuum assumption becomes invalid.33 Jahnisch and co-authors3 observed the laminar flow for Re numbers up to 2000. However, many researchers realized turbulent flows for much lower Re numbers. Turbulence effects become very important for Reynolds numbers above 1000, mainly due to (upstream) geometric non-uniformities.29 It was shown that the channel aspect ratio and the angle of merging of two inlet channels substantially influence the critical Reynolds number. The results, obtained for smooth glass channels with relative roughness around 1%, shown a fast decrease of the critical Reynolds number (from around 2000 at aspect ratio of 2), until the aspect ratio reaches a value of 6 (critical Reynolds number is around 410).34 At laminar flow conditions, the velocity profile fully developed in the direction of the least dimension ([W,-W]) can be described as a function of y position only (Eq. 1): (1) The reaction formula and the rate equation of an irreversible homogeneous reaction proceeding in the micro-reactor are as follows: a\ + pR -> p?, {-rA)= kc"Ac*B (2) where A and B are the reactants and P product; (-rA) and k are the reaction rate (kmol m-3s-1), and the rate constant (m3(a+p-i k mol-(a+3-1)s-1), respectively; cA and cB are the molar concentrations of the components A and B (kmol m-3s-1); and a and ¡3 are the reaction orders in the kinetics term. Figure 1. Scheme of the microchannel with a parabolic velocity profile developed in the smallest dimension. The mass conservation equations, containing convection, diffusion, and the reaction terms for components A and B at steady-state conditions in a dimensionless form with the associated boundary conditions for 2D partial differential equations are given by: Component A V 'dc PeA d2X d-X + dw3 -Da,X"Y'' (3) b.c. , , 17, 0 <[//<} L J J [0, — 1 m^ ci// W (4) V PeE -l 10) results in the shifting away of the model simulations with the PFR curve. Namely, higher values of Pe mean higher velocities and smaller diffusion coefficients, which is reflected in shorter residence and reactions times. Only in the case of very low velocities, expressed as Pe values below 0.1, the microreactor does shift away from the CSTR behavior (Figure 6). This phenomenon might be explained by the fact that at such low velocities and higher values of diffusion coefficients (e.g., Pe = 0.01 at vmax = 0.0001 ms-1 and DA= 6.6 x 10-7 m2s-1) the diffusion of the components in the flow direction (x-direction) becomes more significant. The numerical results of the convection-diffusion-reaction dynamics in the microchannel for the selected geometry (width/length = 200^m/1cm) based on non-linear reaction kinetics are presented in Figure 7 for a wide range of arbitrary reaction rates orders a and ¡ . Figure 6. The performances of the ideal PFR and CSTR classical macro-reactors in comparison with the microreactor for a wide range of theoretical operating conditions, expressed as different Pe numbers (0.01 < Pe < 10). reactant B Figure 7. Averaged dimensionless outlet concentrations of components A and B as the function of the variety of the reaction rates orders a and 3 (-rA = k X c^; k = 0.1 m3(a+3-1) kmol-(a+3-1)s-1; vmx=0.002 ms-1; DA = 1 X 10-9 m2s-1; DB = 2 x 0-9 m2s-1). In order to assure the stability of the described iterative numerical procedure for solving non-linear systems, simple mathematical manipulation is recommended. Namely, the root-finding algorithm based on Newton's method requires the differentiable function fon the Euclidean «-space. Therefore, we define the power exponents in the reaction terms DajXaYp (Eqs. 3 and 5) as Da/X2)a/2(Y2)^/2. Because the power exponent is usually defined as ea lnX, the proposed mathematical manipulation assures the stability of the numerical iteration in cases of feasible appearances of very small but still negative values of dependent variables X and Y. Some typical simulated graphical results of the enzyme reaction process between two immiscible fluids at steady state conditions (Eqs. 9-12) with the velocity profile of the two-phase fluid within the microreactor are presented in Fig. 8. The results of the numerical simulation of a parallel fluid flow in the microchannel with the position of the interface area in the middle of the channel (Fig. 8.a) revealed that at steady-state conditions a fully developed profile takes place very shortly after the beginning of the microchannel, which was shown also in our previous work on steady-state two-phase extraction in the micro-reactor.39 The real values of model parameters for the selected enzyme reaction-diffusion process were used in the simulations. Molecular diffusion coefficients of acetic acid (B) and isoamyl alcohol (A) in water (w) and in n-he-xane (h) were estimated by the Scheibel empirical correlation (DA^h=5.33 x 10-9m2/s, DB/h = 9.85 x 1Q-9m2/s, DB/w = 1.29 x 10-9m2/s), and the partitioning coefficient KP for acetic acid in ternary system with water and n-hexane was found to be 0.0167.45 The kinetic model parameters were estimated according to the literature and are summarized in Table 1.40 Table 1. Chosen binding and inhibition constants and kinetic parameters used in simulations. vr,max'103 (mol/m3 s) kb (mol/m3) ka (mol/m3) Ki,B (mol/m3) 0.32 0.26-10-3 0.96-10-3 2.00-10-3 The results of numerical simulation of reactant B dimensionless concentration profile in water phase YBW and organic phase YB/h at defined conditions with equimolar starting reactants concentration 0.5 x 10-3 mol/m3 are presented in Fig. 8b and c, respectively. As expected due to the low Kp value of B in organic-water system, the component B readily diffused from the organic phase into the b) d) Figure 8. (a) A 3D presentation of a flow pattern of both phases in a microchannel at fw = 10 pl/min and fh = 34 pl/min using Navier-Stokes equations. Graphical presentation of results of a numerical simulation for YB concentration profile along the microchannel (b) in organic and (c) in aqueous phase at z = 0.5H/W. (d) 3D graphical isosurface presentation of concentration of reactant B YB/w in organic phase. aqueous phase. As seen from Fig. 8b, it was almost completely removed from the organic phase and consequently, the concentration of B in the aqueous phase increased at the beginning of the microchannel and reached 3-times higher values than in organic phase due to approximately 3-fold slower flow rate of aqueous phase regarding the organic phase, while further decrease in acid concentration was due to the reaction at the two-liquid boundary. 3D graphical isosurface presentation of dimensionless concentration of reactant B in organic phase YB/h is presented in Figure 8.d where the evident influence of the velocity distribution on the concentration profile can be observed. The process optimization and the determination of real kinetic parameters are feasible by the developed model and experimental work.27 4. Conclusions The theoretical descriptions and analysis of the convection-diffusion-reaction processes for the irreversible homogeneous and non-homogeneous catalytic reactions in a microreactor were investigated. Numerical experiments of a 2D model, developed for the specific case without a chemical reaction taking place in the microchannel, accurately predict the reasonably expected outlet concentrations of both components. The microreactor model simulations were compared with the ideal plug-flow reactor and continuous stirred-tank reactor predictions for a broad range of operating conditions in order to asses under what conditions (microchannel geometry, diffusion, convection, chemical kinetics) a given microreactor is more efficient or productive than a classical macro-reactor. The comparison of the model simulations with the ideal reactor predictions effectively shows close to PFR behavior at 0.5 < Pe < 10. The proposed numerical procedure based on a relatively simple implicit finite difference technique which is improved by a non-equidistant partition of grid points assures accurate, stable, and theoretically consistent solutions of the non-linear systems. The developed and presented 3D mathematical model comprising of flow distribution, transport phenomena and enzyme reaction kinetics, could easily be applied to different reaction types in homogeneous and heterogeneous systems. Further mi-croreactor design and process optimization by means of the use of the optimized microchannel geometry, reac-tants/enzyme ratio, and the determination of real kinetic parameters are feasible by the developed model and additional experimental work. 5. References 1. K. F. Jensen, Chem. Eng. Sci., 2001, 56, 293-303. 2. O. Worz, K. P. Jackel, T. Richter and A. Wolf, Chem. Eng. Sci., 2001, 56, 1029-1033. 3. K. Jähnisch, V. Hessel, H. Löwe and M. Baerns, Angew Chem Int Ed., 2004, 43, 406-446. 4. D. M. Roberge, L. Ducry, N. Bieler, P. Cretton, B. Zimmermann, Chem. Eng. Techn., 2005, 28 (3), 318-323. 5 V. Hessel, H. Löwe, F. Schönfeld, Chem. Eng. Sci., 2005a, 60, 2479-2501. 6. V. Hessel and H. Löwe, Chem. Eng. Techn., 2005, 28 (3), 267-284. 7. P. Watts and S. J. Haswell, Chem. Eng. Techn., 2005, 28 (3), 290-301. 8. P. L. Urban, D. M. Goodall and N. C. Bruce, Biotechnology Advances, 2006, 24, 42-57. 9. K. Mae, Chem. Eng. Sci., 2007, 62, 4842-4851. 10. A. Pohar and I. Plazl, Chem. Biochem. Eng. Q., 2009, 23 (4) 537-544. 11. S. K. Ajmera, M. W. Losey, K. F. Jensen, M. A. Schmidt, AIChE J, 2001, 47, 1639-1647. 12. K. S. Ajmera, C. Delattre, A. M. Schmidt, K. F. Jensen, Stud. Surf. Sci. Catal., 2003, 145, 97-102. 13. J. Yoshida, A. Nagaki, T. Iwasaki, S. Suga, Chem. Eng. Techn., 2005, 28, 259-266. 14. G. Wiessmeier and D. Honicke, J Micromech. Microeng., 1996, 6, 285-289. 15. I. M. Hsing, R. Srinivasan, M. P. Harold, K. Jensen, M. A. Schmidt, Chem. Eng. Sci., 2000, 55, 3-18. 16. J. C. Schouten, E. V. Rebrov, M. H. J. de Croon, Chimia, 2002, 56, 627-635. 17. G. N. Jovanovic, P. Znidarsic-Plazl, P. Sakrittichai, K. Al-Khaldi, Ind. Eng. Chem. Res., 2005, 44,44, 5099-5106. 18. S. Maehara, M. Taneda and K. Kusakabe, Chem. Eng. Res. Des., 2008, 86, 410-415. 19. C. N. Baroud, F. Okkels, L. Ménétrier, P. Tabeling, Phys. Rev. E, 2003, 67, 60104. 20. A. N. Waghode, N. S. Hanspal, I. M. T. A. Shigidi, V. Nasse-hi, K. Hellgardt, Chem. Eng. Sci., 2005, 60, 5862-5878. 21. J. Atencia and D. J. Beebe, Nature, 2005, 437, 648-655. 22. N. Aoki, S. Hasebe and K. Mae, AIChE J., 2006, 52, 1502-1515. 23. D. Bothea, C. Stemichb and H. J. Warneckeb, Chem. Eng. Sci., 2006, 61, 2479-2501. 24. M. V. Kothare, Computers & Chemical Engineering, 2006, 30, 1725-1734. 25. R. Kieffer, C. Charcosset, F. Puel, D. Mangin, Computers & Chemical Engineering, 2008, 32, 1325-1333. 26. M. Tisma, B. Zelic, D. Vasic-Racki, P. Znidarsic-Plazl and I. Plazl, Chem. Eng. J, 2009, 149, 383-388. 27. P. Znidarsic-Plazl and I. Plazl, Proc. Biochem., 2009, 44, 1115-1121. 28. A. Pohar, P. Znidarsic-Plazl and I. Plazl, Lab Chip, 2009, 9, 23, 3385-3390. 29. J. Koo and C. Kleinstreuer, Journal of Micromechanics and Microengineering, 2003, 13, 568-579. 30. S. V. Gokhale, R. K. Tayal, V. K. Jayaraman, B. D. Kulkarni, International Journal of Chemical Reactor Engineering, 2005, 3, R2. 31. T. Bayraktar and S. B. Pidugu, International Journal of Heat and Mass Transfer, 2006, 49, 815-824. 32. D. Jie, X. Diao, K. B. Cheong, L. K. Yong, Journal of Micro-mechanics and Microengineering, 2000, 3, 372-279. 33. K. A. Alfadhel and M. V. Kothare, Chem. Eng. Sci., 2005, 60, 2911-2926. 34. A. Pohar and I. Plazl, Ind. eng. chem. res., 2008, 47, 74477455. 35. E. L. Paul, V. A. Atiemo-Obeng and S. M. Kresta (Ed.): Handbook of Industrial Mixing: Science and Practice, John Wiley and Sons, Hoboken, New Jersey, 2004, pp. 1187-1246. 36. R. B. Bird, W. E. Stewart and E. N. Lightfoot (Eds.): Transport Phenomena, 2nd ed., John Wiley & Sons, New York, 2001. 37. R. C. Aguirre and H. J. Catrakis, Physics of fluids, 2005, 17, 038103. 38. J. P. G. Villaluenga and Y. Cohen, Journal of Membrane Science, 2005, 260, 119-130. 39. P. Znidarsic-Plazl and I. Plazl, Lab Chip, 2007, 7, 883-889. 40. M. D. Romero, L. Calvo, C. Alba, A. Daneshfar, Journal of Biotechnology, 2007, 127, 269-277. 41. M. J. Berger and P. Collela, J. Comput. Phy., 1989, 82, 64-84. 42. P. B. Bochev, G. Liao, G. C. de la Pena, Numerical Methods for Partial Differential Equations, 1996, 12, 489-506. 43. F. Liu, S. Ji and G. Liao, SIAM Journal on Scientific Computing, 1998, 20, 811-825. 44. X. X. Cai, D. Fleitas, B. Jiang, G. Liao, Computers and Mathematics with Applications, 2004, 48, 1077-1085. 45. R. H. Perry, D. W. Green and J. O. Maloney (Eds.): Perry's Chemical Engineer's Handbook, 7th ed., McGraw-Hill Companies Inc., New York, 1999. Povzetek Delo predstavlja teoretičen opis z numeričnimi eksperimenti in analizo reakcijsko difuzijskih procesov homogenih in nehomogenih reakcij v mikroreaktorju. Razvit matematični model v brezdimenzijski obliki vključuje tudi hitrostni profil za laminarni tok mešljivih in nemešljivih tekočin v mikrokanalu pri stacionarnih pogojih. Rezultati napovedi obnašanja mikroreaktorja so primerjani z izračuni v klasičnem kontinuirnem mešalnem in pretočnem reaktorju za primer nepovratne reakcije dveh vstopnih komponent. Primerjava učinkovitosti različnih tipov reaktorjev je podana za široko območje obratovalnih pogojev, izraženih z Pe številom (0.01 < Pe < 10). Za reševanje kompleksnih nelinearnih sistemov je predlagana metoda končnih razlik, izboljšana z uvedbo neekvidistantnih razlik.