Experimental Characterization of an Ion Transport Membrane (ITM) Reactor for Methane Oxyfuel Combustion by Daniel Jolomi Apo B.Eng. Mechanical Engineering, University of Benin, Nigeria. 2007 Submitted to the Department of Mechanical Engineering in partial fulfillment of the requirements for the degree of ARCHIVES Master of Science in Mechanical Engineering at the MASSACHUSETTS INSTITUTE OF TECHNOLOGY February 2012 @ Massachusetts Institute of Technology 2012. All rights reserved. Author ................ ............................. Department of Mechanical Engineering January 3, 2012 , 1 j, C ertified by ......................... YAV I( Ahmed F. Ghoniem Ronald C. Crane ('72) Professor wisor .T)S /A Accepted by.............................D o David E. Hardt mom Chairman, Department Committee on Graduate Theses 2 Experimental Characterization of an Ion Transport Membrane (ITM) Reactor for Methane Oxyfuel Combustion by Daniel Jolomi Apo Submitted to the Department of Mechanical Engineering on January 3, 2012, in partial fulfillment of the requirements for the degree of Master of Science in Mechanical Engineering Abstract Ion Transport Membranes (ITM) which conduct both electrons and oxygen ions have been investigated experimentally for oxygen separation and fuel (mostly methane) conversion purposes over the last three decades. The fuel conversion investigations typically involve converting methane to syngas or higher hydrocarbons. Over the past decade, ITMs have received considerable interest in the industry for oxygen separation and production of syngas. There is also a possibility that the future of ITM industrial use lies with clean power generation as long as stable ITMs are developed which separate oxygen from air and enable reaction of methane to produce carbon dioxide, steam, and the heat which powers turbines. This would hopefully provide CO 2 capture compatibilities at a lower financial and energetic cost than 'conventional' methods. In this thesis, an analysis of reported experimental ITM reactors is presented, with a view to understanding the processes which govern the permeation of oxygen, conversion of methane and production of desired gas species. The analysis shows that temperature and mass flow influence the oxygen permeation within the reactor. Also, the influence of fuel/0 2 ratio on fuel conversion and CO selectivity is discussed. The design and operation of a novel ITM reactor for the experimental investigation of oxygen separation and oxyfuel combustion (a relatively new ITM concept) is presented. The ITM reactor was designed with the aim of providing insight into the use of ITMs for power generation. The reactor has provisions for optical and spatial analysis. The reactor was used to conduct experiments using a Lao.gCao. 1 FeO 33 (LCF) membrane. The results of experiments conducted are presented to show the effects of temperature, mass transfer, and fuel on oxygen permeation, fuel conversion and species selectivity. A comparison is made between the observed results and reported values in literature. Thesis Supervisor: Ahmed F. Ghoniem Title: Ronald C. Crane ('72) Professor Acknowledgments I would like to thank my advisor, Professor Ahmed Ghoniem, for his guidance during my Masters program. His advice and encouragement always proved invaluable for my success in research. I am grateful to my master's degree research mentor, Dr. Patrick Kirchen, from whom I have gained a lot of research knowledge. I would also like to thank him for doing the background work and design of the reactor on which this thesis is based. I am grateful to my research colleagues in the reacting gas dynamics laboratory and other friends at MIT for all the good times. I'm grateful to Dr. Susumu Imashuku and Dr. Lei Wang of the MIT electrochemical energy laboratory for their help with the initial testing of the membrane on which this thesis is based. I would like to thank my research and fellowship sponsors - KAUST, KFUPM, Ceramatec, and TOTAL. I'm grateful to my parents and siblings. They were my source of strength during difficult times. Finally, I'm grateful to GOD. 6 Contents 1 Introduction 1.1 The CO 2 Problem. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.2 Solutions Proffered . . . . . . . . . . . . . . . . . . . . 1.3 Thesis Focus . . . . . . . . . . . . . . . . . . . . . . . . 19 19 2 State of the Art . . . . . . . . . . . . . . . . . .. 2.1 ITM Background. 2.1.1 Membranes and Oxygen Transport Membranes 2.1.2 Mixed Ionic-Electronic Conducting Membranes 2.1.3 ITM Concepts 2.1.4 Usage history and applications . . . . . . 2.1.5 Fundamentals of Oxygen Flux . . . . . . 2.2 Reactive ITM Applications . . . . . . . . . . . . 2.2.1 Oxidative Coupling of Methane . . . . . 2.2.2 Syngas Production . . . . . . . . . . . . 2.2.3 Oxymel Combustion . . . . . . . . . . . 2.2.4 Comparison of ITM methane applications 2.2.5 Reactive ITM challenges.. . . . . . .. 2.3 Methodology for ITM Reactor Characterization 2.3.1 Sealing and Leak Detection . . . . . . . 2.3.2 Experimental Methodology . . . . . . . . 2.3.3 Methods for Analysis . . . . . . . . . . . 2.4 Analysis of ITM Experimental Investigations . . 2.4.1 Temperature Analysis . . . . . . . . . . 2.4.2 Mass Transfer Analysis . . . . . . . . . . 2.4.3 Reactive Analysis . . . . . . . . . . . . . 2.5 Conclusions . . . . . . . . . . . . . . . . . . . . 25 26 26 29 32 34 37 45 46 47 49 50 51 52 53 55 57 60 65 69 72 74 . 3 . . . . . . . Experimental Approach 3.1 Reactor Design and Installation 3.1.1 Reactor . . . . . . . . . 3.1.2 3.2 Reactor Enclosure. . . . 3.1.3 Reactor Sealing . . . . . Membrane . . . . . . . . . . . . 3.2.1 Membrane Details . . . . . . 89 3.3 3.4 3.5 3.6 3.2.2 Pre-operation Membrane Analysis . . . Reactor Process Control and Instrumentation 3.3.1 Plumbing . . . . . . . . . . . . . . . . 3.3.2 Instrumentation . . . . . . . . . . . . . 3.3.3 Gas Chromatography . . . . . . . . . . 3.3.4 Heating . . . . . . . . . . . . . . . . . 3.3.5 Safety Provisions . . . . . . . . . . . . Experimental Methodology and Procedures . 3.4.1 Overview of experimental procedures 3.4.2 Temperature Control . . . . . . . . . . 3.4.3 Inlet Gas Control . . . . . . . . . . . . Methodology for Analysis . . . . . . . . . . . 3.5.1 Permeation-only Analysis . . . . . . . 3.5.2 Reactive Analysis . . . . . . . . . . . . Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90 92 96 97 103 104 107 110 112 113 113 114 115 117 123 4 Results and Analysis 4.1 Tem perature . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2 M ass Transfer . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.3 Reactive Analysis . . . . . . . . . . . . . . . . . . . . . . . . . 4.3.1 Effect of Fuel Inlet Fraction. . . . . . . . . . . . .. 4.3.2 Oxymel Analysis . . . . . . . . . . . . . . . . . . . . . 4.3.3 Influence of Reactive (reducing) Sweep Gas on Oxygen Flux 4.4 Comparison with other ITM Experimental Investigations . . . 4.4.1 Temperature Comparison. . . . . . . . . . . . . .. 4.4.2 Mass Transfer Comparison . . . . . . . . . . . . . . . . 4.4.3 Reactive Comparison . . . . . . . . . . . . . . . . . . . 4.5 Conclusions...... . . . . . . . . . . . . . . . . . . . . . 125 126 128 130 130 134 135 136 136 138 139 140 5 Conclusions 5.1 Summary ............. 5.2 Outlook and Future Work . . 5.2.1 Global Measurements . 143 143 144 144 146 5.2.2 Spatial analysis . . . . Appendices 157 A Reactor within Enclosure 159 B Methodology for Leak Quantification (under non-permeation conditions) 161 C Flow Procedures 163 List of Figures CO 2 Capture in Power Plant Systems: Comparison of Different Natural Gas Technologies. Adapted from Kvamsdal et. al [1] . . . . . . . . . . . . . . . . 20 Four common membrane concepts for oxygen ion conduction: (a) solid electrolyte membrane, (b) dual phase membrane, (c) mixed ionic-electronic conducting membrane (MIEC), and (d) asymmetric porous membrane with graded porosity. Adapted from Bouwmeester et al. [2] . . . . . . . . . . . . . . . . . 27 .. .... .... ... ..... ... ... 30 32 33 2-18 2-19 2-20 . . . . . . . . . . . . . . . . . Common membrane configurations.. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ITM operation.. . . . AZEP process flow sheet, showing the flux of oxygen in the ITM (or MCM) reactor. Adapted from Griffin et al. [4] . . . . . . . . . . . . . . . . . . . . . Oxygen migration in ion transport membranes. Adapted from Liu et al. [5] Important sections in oxygen permeation through ion transport membranes. pO2 is the chemical potential of 02. Adapted from Sunarso et al. [6] . . . . . Effect of membrane thickness on the limiting step during oxygen permeation. Adapted from Sunarso et al. [6] . . . . . . . . . . . . . . . . . . . . . . . . . Surface reaction using a catalyst layer [7] . . . . . . . . . . . . . . . . . . . . Reactions during oxidative coupling of methane [8] . . . . . . . . . . . . . . Production of syngas using an ITM reactor [9] . . . . . . . . . . . . . . . . . Reported effects of temperature on oxygen permeation flux for: (a) separationonly cases; and (b) reactive cases. . . . . . . . . . . . . . . . . . . . . . . . . Ratio of reactive/non-reactive pre-exponential factors . . . . . . . . . . . . . Normalized fluxes as a function of temperature for: (a) separation-only cases; and (b) reactive cases. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Reported effects of Temperature on CH 4 conversion . . . . . . . . . . . . . . Non-reactive oxygen flux dependence on Reynolds number . . . . . . . . . . Reactive oxygen flux dependence on: (a) Reynolds number; and (b) residence time. The sweep gas is a mixture of an inert gas (He or Ar) and CH 4 . . . . . Dependence of CH 4 conversion on sweep mass flow rate . . . . . . . . . . . . Dependence of CH 4 conversion on fuel/0 2 ratio. . . . . . . . . . . . . . . . . Dependence of CO selectivity on fuel/0 2 ratio. . . . . . . . . . . . . . . . . . 3-1 3-2 3-3 Basic schematic of the ITM reactor . . . . . . . . . . . . . . . . . . . . . . . The ITM Reactor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Reactor front cross-section.. . . . . . . . . 79 80 81 1-1 2-1 2-2 An ideal perovskite structure 2-3 2-4 2-5 2-6 2-7 2-8 2-9 2-10 2-11 2-12 2-13 2-14 2-15 2-16 2-17 [3]. 36 38 39 40 46 46 48 66 67 68 69 70 71 71 73 74 3-4 3-5 3-6 3-7 3-8 3-9 3-10 3-11 3-12 3-13 3-14 3-15 3-16 3-17 3-18 3-19 3-20 3-21 3-22 3-23 3-24 3-25 4-1 The ITM Reactor within its insulation.. . . . . . . . . . . . . . . . . . . The different sealants used for the ITM reactor . . . . . . . . . . . . . . . . Sealant performance tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . Sealant performance results: quantified in pmol.s- 1 of oxygen (from the air which leaks into the reactor sweep side) . . . . . . . . . . . . . . . . . . . . . Comparison of total leak at 500sccm air and 500sccm C0 2 , to the expected oxygen permeation flux (assumed to be in the order of 1umol.cm- 2 .s- 1 ) . . . . Reactor top cross-section . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Pre-operation EDX graph . . . . . . . . . . . . . . . . . . . . . . . . . . . . Pre-operation XRD graph . . . . . . . . . . . . . . . . . . . . . . . . . . . . Pre-operation SEM image . . . . . . . . . . . . . . . . . . . . . . . . . . . . Reactor Process Control and Instrumentation (PC&I) . . . . . . . . . . . . . Measurement locations on the ITM reactor . . . . . . . . . . . . . . . . . . . Reactor Plumbing Schematic . . . . . . . . . . . . . . . . . . . . . . . . . . . The major reactor instrumentation . . . . . . . . . . . . . . . . . . . . . . . Reactor thermocouples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Membrane temperature measurement configuration . . . . . . . . . . . . . . Pyrometer calibration data: (a) membrane temperature vs pyrometer voltage output; (b) membrane emissivity and pyrometer current output vs membrane temperature (E= (62.5 * I + 323) /T) . . . . . . . . . . . . . . . . . . . . . . Oxygen sensor calibration: (a) calibration duct; (b) calibration data . . . . . Reactor Gas Chromatography . . . . . . . . . . . . . . . . . . . . . . . . . . Reactor Heating Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . The Enclosure Heaters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Heating tests with the cartridge and enclosure heaters (SP = Set Point). N.B: The green vertical lines indicate the cartridge heater locations . . . . . . . . Reactor Safety Control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82 84 86 86 87 90 91 91 92 93 94 97 97 98 99 100 102 104 105 106 107 108 Dependence of oxygen permeation flux on membrane temperature (non-reactive). Feed (air) flow = 500sccm, sweep (C0 2) flow = 500sccm. . . . . . . . . . . . 126 4-2 Bulk temperatures within reactor feed and sweep sides . . . . . . . . . . . . 127 4-3 Dependence of oxygen permeation flux on Sweep (C0 2 ) flow (non-reactive). Membrane temperature = 800'C, feed (air) flow = 500sccm. . . . . . . . . . 128 4-4 Dependence of oxygen permeation flux and sweep 02 partial pressure on the sweep Reynolds number (non-reactive). Membrane temperature = 800'C, feed (air) flow = 500sccm. Re = (p#) / (pzDc), De = 2.24cm. . . . . . . . . . . . 129 4-5 Dependence of (a) oxygen permeation flux; (b) sweep oxygen partial pressure, and fuel/0 2 ratio; on fuel inlet concentration. Membrane temperature = 800 0 C, CO 2 inflow = 500sccm, feed (air) flow = 500sccm. . . . . . . . . . . . 130 4-6 Dependence of fuel conversion and species selectivities on fuel inlet concentration. Membrane temperature = 800 0 C, CO 2 inflow = 500sccm, feed (air) flow = 500secm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131 4-7 Possible interaction of fuel with permeated oxygen in the reactor sweep side 132 4-8 4-9 4-10 4-11 4-12 4-13 Analysis of oxymel products obtained from reaction between fuel and 02 in the ITM reactor. Membrane temperature = 800'C, CO 2 inflow = 500sccm, feed (air) flow = 500sccm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Influence of fuel addition on oxygen permeation flux in the ITM reactor. Membrane temperature = 800'C, feed (air) flow = 500sccm, sweep flow = 500sccm CO 2 + variable CH 4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Oxygen flux dependence on temperature (non-reactive) for the ITM reactor compared with reported investigations. . . . . . . . . . . . . . . . . . . . . . Normalized oxygen flux dependence on temperature (non-reactive) for the ITM reactor compared with reported investigations. See section 2.4.1 for . . . . . . . . . . . normalization methodology.. . . . . . . . . . . . . . Normalized flux dependence on sweep mass flow (non-reactive) for the ITM reactor compared with reported investigations. See section 2.4.1 for normal. . . . . . . . . . . . . . . . ization methodology.. . . . . . . . . . . . . Normalized flux dependence on Residence time (non-reactive) for the ITM 134 135 137 137 138 reactor compared with reported investigations. rres = V/V. See section 2.4.1 for normalization methodology. . . . . . . . . . . . . . . . . . . . . . . . . . 139 4-14 Dependence of CH 4 conversion on fuel/0 2 ratio for the ITM reactor compared with reported investigations. . . . . . . . . . . . . . . . . . . . . . . . . . . . 140 A-1 The ITM reactor within its enclosure in the laboratory. . . . . . . . . . . . . 160 C-1 Reactor flow diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163 12 List of Tables 2.2 Comparison between ITM investigations in oxidative coupling of methane (OCM), syngas production and oxymel combustion. . . . . . . . . . . . . . . Experimental investigations from literature considered for analysis . . . . . . 3.1 3.2 Process Control Equipment and Instrumentation . . . . . . . . . . . . . . . . 95 . . . . . . . . . . . . . . . . . . . . 112 ITM reactor measurements.. . . . . 4.1 4.2 Experimental points considered for analysis . . . . . . . . . . . . . . . . . . . 126 Mass flow rates at different planes below the membrane (inlet sweep velocity = lcm/s). Results obtained from 2-D numerical analysis of the reactor . . . 133 5.1 Experimental points for future consideration . . . . . . . . . . . . . . . . . . 145 C.1 Flow procedures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164 2.1 50 62 14 Nomenclature Abbreviations DAQ data acquisition system, page 94 EDX energy-dispersive x-ray spectroscopy, page 57 GC gas chromatograph, page 103 ITM ion transport membrane, page 21 MIEC mixed ionic-electronic conductor, page 26 0CM oxidative coupling of methane, page 47 PC&I process control and instrumentation, page 22 POM partial oxidation of methane, page 47 sccm standard cubic centimeters per minute, page 80 SEM scanning electron microscopy, page 57 XRD x-ray diffraction, page 57 Symbols uhi ni i mass flow rate of gas i, page 96 molar flow rate of gas i, see equation (3.11), page 117 volume flow rate of gas i , see equation (2.12), page 54 A membrane surface area exposed to gases on one side, page 117 AO pre-exponential factor for flux normalization, see equation (2.20), page 65 D* tracer diffusion coefficient, see equation (2.2), page 40 De characteristic reactor length, see equation (2.22), page 70 Ds self-diffusion coefficient, see equation (2.2), page 40 Dv vacancy diffusion coefficient, see equation (2.8), page 44 Ea activation energy, see equation (2.20), page 65 F Faraday constant, see equation (2.3), page 41 h* electron hole, page 38 balanced surface exchange rate at equilibrium, see equation (2.7), page 43 e J 2 normalized oxygen permeation flux, see equation (2.21), page 66 Jo 2 oxygen flux, page 33 k, forward surface reaction rate constant, see equation (2.8), page 44 kr reverse surface reaction rate constant, see equation (2.8), page 44 ks surface exchange coefficient, see equation (2.2), page 40 L membrane thickness, page 40 Le characteristic thickness, page 34 MA molar mass of gas i, page 115 mfj mass fraction of gas i, page 116 n number of carbon atoms in the molecule of a carbon-containing product i , see equation (2.19), page 60 00 lattice oxygen, page 38 P0 low oxygen partial pressure, page 28 high oxygen partial pressure, page 28 R universal gas constant, see equation (2.3), page 41 Rm log-mean radius, see equation (2.9), page 44 Ro outer radius, see equation (2.9), page 44 Rin inner radius, see equation (2.9), page 44 Re Reynolds number, see equation (2.22), page 70 S, selectivity of carbon-containing product p , see equation (2.0), page 34 T membrane temperature, see equation (2.3), page 41 t el electronic transfer number, see equation (2.6), page 43 16 tel electronic transference number, see equation (2.1), page 37 tion ionic transfer number, see equation (2.6), page 43 tion ionic transference number, see equation (2.1), page 37 V volume, see equation (2.23), page 70 V** oxygen vacancy, page 38 Xi molar fraction of gas i, see equation (2.15), page 58 Xr conversion of reactant gas r, see equation (2.0), page 34 Y yield of carbon-containing product p , see equation (2.0), page 34 C2 organic compound containing two carbon atoms, page 34 e- electron, page 26 Greek p viscosity, see equation (2.22), page 70 p density, see equation (2.22), page 70 o-e_ partial electronic conductivity, see equation (2.3), page 41 o-ion partial ionic conductivity, see equation (2.3), page 41 o-total total ionic conductivity, see equation (2.6), page 43 7- timescale, see equation (2.23), page 70 Subscript in gas flows into the reactor, see equation (2.15), page 58 leak leakage of gas into reactor sweep side, see equation (2.12), page 54 meas gas concentration measured at sweep side exit by gas chromatograph or oxygen sensor, page 118 out gas flows out of the reactor, see equation (2.15), page 58 perm gas permeation across membrane from feed side to sweep side, page 120 s property of reactor sweep side, see equation (2.15), page 58 18 Chapter 1 Introduction 1.1 The CO 2 Problem The general concern about carbon dioxide (C0 2 ) levels in the atmosphere is in its ability to make the earth warmer. Climate model studies indicate that temperatures everywhere on earth would increase by 1.5 - 4.5'C if the concentration of CO 2 is doubled in the atmosphere. The increase in temperatures globally appear to support the results of these studies. Even though some skeptics remain doubtful of the influence of CO 2 on global warming, it is commonly agreed that further increases in CO 2 emissions and global warming need to be forestalled [10]. One important source of CO 2 emission is natural gas (95% methane) combustion, and it is currently a very important energy resource around the world. According to the United States Energy Information Administration, natural gas contributed 5.93% of the world's energy-related carbon dioxide emissions in 2007. The projection for natural gas contribution in 2035 is 8.59% [11]. Furthermore, it is expected in the future that the use of natural gas will 19 be limited by rigorous emissions policies, unless carbon capture and sequestration provide a favorable competition with low-carbon energy sources [12]. Therefore, studies into CO 2 emissions reduction would help provide efficient and clean ways to use natural gas (especially in power plants), and go a long way to help reduce global warming and aid the capture and reuse of carbon dioxide. 1.2 Solutions Proffered There are several technologies currently used in natural gas power plants to facilitate carbon capture, as shown in figure 1-1. Pre-capture, post-capture, and oxymel techniques are currently being used even though the technologies for their implementation are relatively new. ATR MSR-H 2 Autothermal reforming + water-gas shift Water cycle Chemical looping combustion Membrane H2 separation AZEP Advanced Zero Emission WC CLC 50.0% - - 4- * *47 % 47 % 45 45.0% 40.0% o L di 35.0% 0_0 X_g# C~i , (ITM) Process 100% CO2 Capture -5%C02 Capture SOFC GT Solid Oxide Fuel Cell Gas Turbine **Coo .* ITM: Ion Transport Membrane Figure 1-1: CO 2 Capture in Power Plant Systems: Comparison of Different Natural Gas Technologies. Adapted from Kvamsdal et. al [1] Pre-capture involves the removal of CO 2 after carrying out steam reforming of fuel before 20 the combustion process. Post-capture involves the removal of CO 2 after the combustion process. Oxymel techniques involve the combustion of fuel and oxygen. For the pre-capture and post-capture cases however, separating CO 2 from other pollutants using solvents remains problematic considering the overall efficiencies of these plants. Oxymel combustion is the best method to burn natural gas cleanly. If the equivalence ratio during combustion is 1, the products of combustion would ideally be carbon dioxide (C02 ) and water (H2 0), and carbon dioxide can easily be captured after condensation of the water. However, the costs of obtaining oxygen by cryogenic distillation or adsorption techniques are generally high. For example, cryogenic distillation facilities account for 50% of the construction costs of some syngas plants [13,14]. A novel approach is to use ion transport membranes (ITM) for separation of oxygen from air before combustion with natural gas (e.g AZEP in Fig 1-1), with efficiency projections currently above 50%. ITMs are metallic oxide membranes permeable only to one gas specie (oxygen in this case) when subjected to high temperatures (> 700) and a partial pressure gradient of the gas. The use of ITMs for separation of oxygen from air, and combustion with fuels are not new concepts. However, their use for oxymel combustion is relatively unexplored. Two limitations currently reported with the use of ITM reactors for oxymel combustion are [15]: " formation of C 2 4 , C2H, and CO, making CO 2 capture difficult, and * low oxygen flux and methane conversion in the absence of a catalyst. In addition to the above limitations, there is a distinct lack of reports on spatial analysis or optical analysis of the reaction zone within ITM reactors which would aid the study of different gradients and species formation in the reaction zone. It is therefore necessary 21 to explore the use of ITM reactors for separation of oxygen from air, complete oxidation of natural gas, as well as the various factors that influence the effective operation of ITM reactors macroscopically and spatially. 1.3 Thesis Focus This report covers the design and development of an ITM reactor, and the use of said reactor for experimental investigation of oxygen separation from air and methane oxymel combustion. The experimental approach (see chapter 3) covers: 1. Design of an ITM reactor with stagnation flow configuration and optical access (to aid spatial and reaction zone analyses), 2. Design of a complete process control and instrumentation (PC&I) system for the reactor, and 3. Development of reactor-centric procedures and equations to aid understanding of the reactions and processes observed. The goals for this work are: 1. Fundamental investigation into the effects of operating parameters (temperatures, inlet gas flow rates, and fuel/diluent ratio) on oxygen flux and the macroscopic thermochemical processes (e.g. yield of reaction products). 2. Investigation into the use of the ITM reactor for methane oxymel combustion. The oxygen flux and the extent of the different reactions (oxidative coupling, syngas pro22 duction, or oxymel) observed will provide an insight into the suitability of ITM reactors for methane oxymel combustion. Oxidative coupling produces C2 H4 and C 2H6 ; syngas production is the partial oxidation of methane into CO and H2 ; and oxymel combustion produces CO 2 and H2 0. It should be noted that the overall focus/aim of this work is not to provide a detailed insight into the use of ITM reactors for power generation or any other full scale application, but to characterize the ITM reactor at a fundamental level. Chapter 2 provides a review of the major concepts regarding ITMs as well as insight into their current usage. An overview of current ITM reactor applications and experimental methodologies is presented. The two major ITM methane applications are oxidative coupling and syngas production. Furthermore, an analysis of reported ITM experimental investigations is conducted. The reported data from the investigations are analyzed to understand some of the underlying governing processes for ITM reactor operation. In chapter 3, the experimental approach for the ITM reactor is explained. The design of the ITM reactor is presented along with an explanation of the membrane used. The design of the reactor PC&I system is explained. Reactor-centric experimental procedures and analysis equations are also presented. An analysis of experimental results is presented in chapter 4 with a view to understanding the current reactor operating regimes and to provide insight for reactor optimization. The effects of membrane temperature and inlet mass flow rate are analyzed based on their impacts on oxygen permeation. Also, an analysis of the effects of fuel use in the reactor is conducted. The main focus in the reactive analysis is on oxygen permeation, fuel conversion, and overall 23 CO 2 yield. Chapter 5 provides a summary of the contribution of this work and recommendations for future research. Chapter 2 State of the Art The use of ion transport membrane (ITM) reactors for oxygen separation and methane combustion is a novel approach, the main benefit being low cost of oxygen separation at relatively high power plant efficiencies. The approach also allows for a wide range of applications. These include the formation of ethylene via oxidative coupling (for polyethylene formation process), production of syngas via partial oxidation (e.g. for the fischer-tropsch process), and formation of carbon dioxide, water vapor and heat via complete oxymel combustion (for carbon capture and heat supply to turbines in power plants). Although the focus of this investigation is oxymel combustion, other modes of ITM operation (oxidative coupling and syngas production) are also discussed here as they provide the majority of the current ITM reactor literature. As seen earlier from Figure 1-1, the efficiencies that can be derived from applying ITM technology to oxymel combustion and power generation are promising compared to other carbon capture technologies. However, it is necessary to understand the concept of an ITM, the characteristics of currently used ITMs, the ITM governing processes, and the principles 25 of oxygen permeation. It is also important to understand the effects of operating parameters on oxygen permeation and the formation of combustion products in order to optimize the ITM reactor and aid complete capture of CO2- 2.1 ITM Background Ion transport membranes (ITMs) are fabricated from ionic and mixed-conducting ceramic oxides that conduct oxygen ions (02-) and electrons (e-) at elevated temperatures. What follows, is a background discussion of the different oxygen transport membrane types, mixed ionic-electronic conducting membranes (abbreviated as MIEC, but also called ITM), ITM concepts and applications, and the governing equations and fundamentals of oxygen flux. 2.1.1 Membranes and Oxygen Transport Membranes In this report, the term 'membrane' refers to a layer of material between two gas streams which allows a gas (e.g. oxygen) to permeate across it due to a driving force. The driving force is the difference in the partial pressures of the permeating gas, or the electric potential gradient across the membrane. Figure 2-1 shows four common concepts for membranes permeable to oxygen. They are [2]: (a) solid electrolyte membrane, (b) dual phase membrane, (c) mixed ionic-electronic conducting membrane (MIEC), and (d) asymmetric porous membrane with graded porosity. Solid electrolyte membrane systems involve sandwiching a solid oxide electrolyte between two gas-permeable electrically conductive electrodes (Figure 2-la). This configuration incorporates the diffusion of oxygen ions into the electrolyte, and catalytic reaction with fuel on cathode node e e (b) A (a) P'o2 02 - 2- e <- P"02 (d) (c) Figure 2-1: Four common membrane concepts for oxygen ion conduction: (a) solid electrolyte membrane, (b) dual phase membrane, (c) mixed ionic-electronic conducting membrane (MIEC), and (d) asymmetric porous membrane with graded porosity. Adapted from Bouwmeester et al. [2] the anode/cathode interface, while giving off water vapor, carbon dioxide, heat, and electrons. The electrons transport from the anode through the external circuit and back to the cathode, providing a source of useful electrical energy in an external circuit (i.e. oxygen pump). Solid electrolyte membranes can produce oxygen at high pressures. This provides an advantage for high pressure applications, since compressors will not be needed [16]. A dual phase membrane, shown in figure 2-1b, is regarded as a metallic phase distribution into the cloud or matrix of an oxygen ion-conducting material/oxide. For example, yttria-stabilized zirconia (YSZ) can be doped with titania [17, 18], ceria [19, 20], or palladium [21]. An increase in dopant concentration reportedly leads to an increase in the electronic conductivity of dual phase membranes [2]. However, the range of solid solubility 27 of these multivalent membranes may limit the increase in electronic conductivity. These membranes broaden the dense ceramic membrane spectrum by adding the solid oxide electrolyte alternative. However, without the application of internal or external circuitry, most dual phase membranes cannot be used practically since they exhibit very low oxygen flux when operated within commonly used ranges of temperature (< 1000C) and oxygen partial pressure (< 0.21 atm). Also, continuous electronic conduction with accompanying oxygen permeability can be damaging to the membrane as evidenced in their applications in fuel cells and oxygen sensors [22]. A mixed ionic-electronic conducting membrane (MIEC) is depicted in figure 2-1c. MIECs typically conduct both oxygen ions and electrons thereby eliminating the need for electrodes or external circuitry. They are dense and impermeable to gas molecules, and allow only oxygen ions and electrons permeate across them. The chemical potential is the driving force for the overall oxygen flux across the membrane. Oxygen dissociates and ionizes at the high pressure membrane surface (feed side, P 2 ) by associating with electrons in accessible surface electronic states. Electronic charge carriers flow simultaneously to compensate for the flux of oxygen ions within the membrane. When oxygen ions reach the low pressure surface (sweep side, P02 ), they recombine and release electrons to form oxygen molecules. The oxygen molecules on the sweep surface then flow into the sweep stream [2]. MIECs are very attractive for practical applications since they conduct both oxygen ions and electrons without the use of external circuitry. Figure 2-1d shows an asymmetric porous membrane with graded porosity. This is usually made of materials such as a-A12 0 3 , 7-Al 2 03, TiO 2 and SiO 2 . Its main advantages are [23] 28 (a) high surface area to volume ratio, and (b) enhanced mechanical support. 2.1.2 Mixed Ionic-Electronic Conducting Membranes As discussed in section 2.1.1, MIECs conduct oxygen ions and electrons without the use of external circuitry. The most commonly used MIECs have been shown to exhibit either a perovskite structure (empirical formula ABO 3 ) or fluorite structure (empirical formula A0 2 , e.g. ZrO 2 ) [6]. A & B are cations: the A-site is usually occupied by rare earth materials while the B-site contains transition metals with mixed valency [24]. Perovskites are used more in practical applications because of their high oxygen fluxes. A MIEC is termed a perovskite if it has a similar crystal structure to calcium titanium oxide (CaTiO 3 , also called the 'perovskite structure') [25]. The name 'perovskite' was coined in honor of Count Lev Alekseevich Perovskii (1792 - 1856), a Russian mineralogist from St. Petersburg, Russia [26]. It should be noted that in the rest of this report, the term 'ion transport membrane' (ITM) is used to refer to MIECs. Composition Perovskites may contain a mixture of two or more elements in the A-site and/or the B-site, e.g. Bao.5 Sro.5 Co0o.Fe0 .2O36 . One version of a perovskite structure is shown in figure 2-2. The A-cations combine with oxygen ions to form a lattice structure with octahedral interstitial sites occupied by the B-cations [27]. The structure is a octahedra with oxygen ions linked to all corners. The A-cations fill the dodecahedral holes while the B-cation fills the octahedral holes. The coordination number of the A-cation is 12, while that of the B-cation is 6. The 29 coordination number of an ion in a crystal is the total number of ions it holds as its nearest neighbors [28]. Active Top x x F Right @02. *A *B Figure 2-2: An ideal perovskite structure [3]. Perovskite membranes have been shown to exhibit higher oxygen fluxes than flourites, even though they are structurally less stable [6]. Teraoka et al. [29-31] showed that oxygen permeation fluxes through LaCoO 3-based perovskite-type membranes were 2-4 orders of magnitude higher than those of stabilized zirconia (a fluorite) at the same temperatures. ITMs which are made to take advantage of the best traits of perovskites and fluorites, have been reported. An example is 40 vol.% Lao.sSro 2 MnO 3 - mixed with 60 vol.% yttria-stabilized zirconia (YSZ) [23]. The metals in perovskite structures can be substituted to enhance performance. It is important to note that the B-site cation improves the oxidation catalytic activity of per30 ovskites [32]. Lin et al. [33] showed that by substituting Fe with Co, Lao.2 Sro.8 CoO 3-3 (LSC) showed better C 2 selectivity than SrCo 0 .sFeO.2 0 3 _o (SCF) during oxidative coupling of methane (OCM). Although A-site cations are generally inactive catalytically, they strongly affect the oxygen non-stoichiometry of the oxide and the flux of oxygen ions across the membrane [34]. I.e. A-site cations influence the oxygen flux while the B-side Cations influence the ITM's catalytic properties. Configurations There are three broad MIEC configurations used in reactors [35] - planar/disk, tubular, and hollow fiber, as shown in Figure 2-3. However, commercial large scale applications utilize multi-planar or multi-tubular MIEC reactor configurations [8]. Hollow fiber membranes in particular, are becoming popular due to their much larger membrane surface area to equipment volume ratio than planar/disc and tubular membranes [36]. However, hollow fiber membranes have less thickness and thus lower mechanical strength compared to tubular membranes leading to difficulties during assembly into membrane reactors [35]. Design variations for porous layers and asymmetric membranes are common. Typically, porous layers are added to membranes to improve the bending strength and surface reactions while catalysts can be added to on any given surface to either improve oxygen flux or improve the results of a desired reaction. Also, the presence of porous layers allows the use of much thinner dense membranes designed to enhance oxygen flux. In the case of tubular or hollow fiber membranes, the inside of the tubes are usually packed with a catalyst to improve desired reactions in the case of a reactive ITM [37,38]. 31 feed side feed out feed i planar/disk cross-section 02 flux s-swi -sweepin sweep out e de diameter 210mm thickness 0.5mm) (a) planar/disk feedUnt O flux, feed side outlet 'sweep side Ne717 Neep et sweep outlet internal dia.> o.1smm t*ickness>0.2Smm (b)tubular feed inlet 02 flux---, tubular cross-section feed side d outlet (c)hollow fiber sweep outlet hollowfiber cross-section 0 internal dia. < 0.15mm thicness< 0.25mm Figure 2-3: Common membrane configurations 2.1.3 ITM Concepts Figure 2-4a shows the application of a two-dimensional planar ITM purely for separation of oxygen from air. High temperatures and high partial pressure gradient of oxygen are required for oxygen flux across the membrane. Under these conditions, some oxygen can be separated from the air on the feed side and permeated to the sweep side where it is 'swept' by a sweep gas with lower partial pressure of oxygen (e.g. C0 2 )In figure 2-4b, the separation process is used in combination with a reactive process. In this case an oxymel reaction occurs in the sweep side resulting in the formation of carbon dioxide and water vapor as products. The use of a reactive sweep gas (e.g. CH 4 ) enhances the oxygen flux beyond the separation-only case. This is believed to be due to the consumption of oxygen in the sweep side during reaction, thereby maintaining a large partial pressure gradient of oxygen across the membrane. Some important terms regarding ITM use for oxygen separation and fuel combustion are: 32 Feed side - high 02 partial pressure (P'o2 ) air M 4cl %W4 o 10CO 4 *1 ITM Temperature Ze700"C l Ni CO 0 2-depleted air CO2 & permeated 02 Sweep side - low 02 partial pressure (P"o2) (a) separation only air $$ $0 Feed side - high 02 partial pressure (P'o2 ) CH4 2-depleted air C02 & H20 CH4 +202 -.. ' C02 + 2H20 Sweep side - low 02 partial pressure (P"o2 ) (b)separation coupled with oxyfuel reaction Figure 2-4: ITM operation. Oxygen flux (J0 2 ): The amount of oxygen passing though the membrane per unit membrane surface area and time [39]. Oxygen partial pressure: This is equivalent to the mole fraction of oxygen in a gaseous mixture. For ITMs, P denotes the higher oxygen partial pressure (on the feed side), while P02 represents the lower oxygen partial pressure (on the sweep side) Characteristic thickness (Lc): The membrane thickness at which oxygen permeation is equally determined by the surface-exchange kinetics and bulk-diffusion [6]. Surface exchange refers to the reactions between oxygen molecules or ions and electrons on the membrane surface while bulk diffusion refers to the transport of oxygen ions within 33 the membrane. Conversion (X): The conversion of a reactant r (e.g. 02, CH 4 ) is the percentage of the total supply of reactant r (total permeated in the case of 02) that is converted to products during a reaction. Selectivity (S): The selectivity of a carbon containing product p (e.g CO, C0 2 , C2 H4 , C2 H) is the percentage of converted fuel (e.g. CH 4 ) that forms product p [40]. Yield (Y): The yield of a carbon containing product p, is the percentage of total methane that forms product p. It is the product of the fuel conversion and the selectivity of p, i.e. Y = Sp x X, [40]. C 2 : This notation refers to products with two carbon atoms (mainly C 2H4 and C 2 H6 ). Feed side: This is the side of the membrane that is exposed to feed gas. Sweep side: This is the side of the membrane that is exposed to permeated oxygen, sweep gas and/or fuel. 2.1.4 Usage history and applications Since the mid 1960s, when rare earth aluminates were first reported to exhibit elevated levels of oxide ion conductivity [41, 42], the development of oxygen-deficient mixed ionicelectronic conducting membranes has become a popular trend [43]. Teraoka et al. [29-31] (1985 - 1991) reported the use of cobalt-rich perovskites with very high oxygen fluxes. These perovskites exhibited high anion deficiencies when subjected to decreased oxygen partial pressures and increased temperatures. By mid 1990s, extensive research in the development 34 of acceptor-doped perovskites with the general formula LajxAxCoiyByO 3 _6 (A = Sr, Ba, Ca and B = Fe, Cu, Ni) were widely common [2]. From mid 1990s, two trends began to emerge in the use of ITMs for fuel conversion purposes. The earliest trend was oxidative coupling for the conversion of methane and ethane to ethylene. Ten Elshof et al. [44] carried out oxidative coupling of methane (OCM) to form ethylene between 800-900C using a Lao0 6 Sr. 4 Co0 .Feo.20 furnace. 3 membrane in a quartz Lin et al. [40, 45, 46] carried out oxidative coupling of methane and ethane us- ing Lao.sSro. 2 Co Fe 0 .4 0 3 _ 3 (LSCF) and Bii.5 Y0 .3 Sm 0 .20 3 (BYS) membranes respectively be- tween 750 - 950C. The methane experiments gave C2 (ethane and ethylene) yields between 10 - 18% while the ethane experiments gave an ethylene yield of 56%. According to Lu et al. [37], a similar experiment using the barium based membrane BaCeo.sGdo.203_6 gave 16.5% C2. The second trend was the partial oxidation of methane to syngas (CO + H2 ). Balachandran et al. [47] carried out the partial oxidation of methane to syngas using La-Sr-Fe-Co-O membrane tubes (each 30cm long) in a reactor at 850'C, obtaining 99% methane conversion. Ikeguchi et al. [48] reports the use of a disk-type SmO4 Bao.6 Co 0 .2Feo.0 O 3-3 membrane for partial oxidation of methane to syngas, with 98% CO selectivity being observed at 900C. Recently, similar experiments were reported by Wang et al. [49] and Takahashi et al. [13] using membranes made of BaCoFeyZrO 3 _o (BCFZ, x + y + z = 1) and La_,SrTijyFeyO3-6 (LSTF, x = 0.1-0.8, y = 0.6-1.0) respectively. BCFZ gave 95.8% CH 4 conversion and 96.6% CO selectivity at 925C, while LSTF gave 60% CH 4 conversion and 99.9% CO selectivity at 9000 C. Of high importance is the work reported by Tan et al. [15]. 35 Apart from working on oxidative coupling of methane [50] and syngas production [35], methane oxymel experiments were carried out using four hollow fiber Lao. 6Sro. 4 Co0 .2 Fe 0.sO3 6 (LSCF6428) membranes in a tubular reactor. The aim was to provide an insight into ITM use for oxycombustion based heat/power generation. Over the past 10 years, the most commonly reported ITMs have been LSCF [51-54] and BSCF [55-57]. BSCF exhibits the highest oxygen flux generally reported, but is mostly limited to separation-only applications because of its chemical and structural instability in reducing environments [58,59]. LSCF is the most widely used for reactive ITM applications because of its good flux and acceptable chemical and structural stability [39]. The most commonly used fuel for ITM reactors is methane (CH 4 ). However, the use of ethane as fuel has also been reported [60]. ITM reactor Steam turbine Bleed gas HX Combustor: Cooling Water Natural gsCO2 compressor CO2 H20 H20 ........ ......... ..................... HRSG Air ...... .... CO2 Oxygen depleted air s.... Gas turbine Figure 2-5: AZEP process flow sheet, showing the flux of oxygen in the ITM (or MCM) reactor. Adapted from Griffin et al. [4] Today, ITM use has evolved to a large extent due to the need for less expensive techniques for oxygen separation in power plants and engines [61]. There are now many different ITM 36 configurations, alternative composite ITM materials [62] and documented industrial applications such as the Zero Emissions Ion Transport Membrane Oxygen Power (ZEITMOP) [63], Zero Emissions Membrane Piston Engine System (ZEMPES) [51,64,65], and Advanced Zero Emissions Power Plant (AZEP) cycles [4,66]. Figure 2-5 shows an AZEP process flow sheet. It can be seen that air is compressed, oxygen is separated from the air by a membrane, and the separated oxygen is reacted with natural gas. The oxygen depleted air (expanded after separation) and flue gas are passed through a heat recovery steam generator (HRSG) to help drive a steam turbine. 2.1.5 Fundamentals of Oxygen Flux The primary purpose of an ion transport membrane is to separate oxygen from a feed gas (e.g. air). It is therefore important to understand the processes that drive the oxygen flux across the membrane in order to properly design and improve membrane performance. The flux of oxygen across an ITM involves the flux of oxygen ions which are charge compensated by a simultaneous flux of electrons in the opposite direction. At any total conductivity, the maximum oxygen flux is said to occur when the transference numbers (ionic and electronic) are the same, i.e. tion = tei= 0.5 (2.1) The transference number of an ion or electron is the fraction of the total current that is carried by that ion or electron in passing through an electrolytic solution or solid [2]. 37 ITM High 02 partial pressure side Oo + 2h' 00000 (P02 ) (4) O' 4.O 1 02 + V3* == O 0=>+2h' 02 + V' (5) -- +02 Sweep |gas 2(2) 02 -- > Feed gas (P0 2 ) Low 02 partial pressure side Figure 2-6: Oxygen migration in ion transport membranes. Adapted from Liu et al. [5] Figure 2-6 shows a schematic diagram describing the mechanism for oxygen permeation through an ion transport membrane. Krbger-Vink notation is used to define the charged defects. 00 stands for lattice oxygen, V ' for oxygen vacancy, and h* for positive electron hole. The process of oxygen permeation from the feed side (high PO2 ) to the sweep side (low P0 2 involves a series of steps: [5]: 1. Mass transfer of oxygen gas from the feed gas stream to the surface of the membrane (high P' 2 side); 2. Reaction on the surface of the membrane (high Po2 side) between oxygen vacancies and oxygen molecules. 3. Bulk diffusion of oxygen ions by membrane vacancies; 4. Reaction on the surface of the membrane (low P02 side) between electron holes and oxygen ions; and 5. Mass transfer of oxygen gas from the surface of the membrane to the sweep gas stream (low P , side). Generally, the resistances between the gas phase and membrane are usually small and negligible. Only membrane bulk diffusion (step 3) and surface exchanges (steps 2 and 4) are taken into consideration in developing transport equations [5,67]. The surface exchanges involve a series of sub processes as explained by [68]. Figure 2-7 shows the important oxygen permeation steps. It can be seen that the chemical potential of oxygen drops from the feed side to the permeate side. This drop in chemical potential is the main driving force for oxygen flux. Surface Exchange zonel Bulk Diffusion 02 Surface Exchange zonell 2 Feed side Sweep side Figure 2-7: Important sections in oxygen permeation through ion transport membranes. p10 2 is the chemical potential of 02. Adapted from Sunarso et al. [6] The overall oxygen permeation rate is expected to be limited by the slowest moving species or the slowest process. However, as the membrane thickness decreases towards the characteristic thickness (i.e. as L -s Lc), the diffusion resistance decreases and the controlling step changes from bulk diffusion to surface kinetics [69-72]. Figure 2-8 shows a schematic 39 of the limiting steps. It should be noted that reducing the membrane thickness will improve oxygen flux. However, once the membrane thickness is reduced below Lc, only marginal (if any) increases in oxygen flux can be obtained [69,70,73]. Kinetics-controlled Diffusion-controlled Lc Membrane thickness Figure 2-8: Effect of membrane thickness on the limiting step during oxygen permeation. Adapted from Sunarso et al. [6] The characteristic thickness (Lc) of an ITM , is the membrane thickness at which the oxygen flux is equally determined by bulk diffusion and surface exchange kinetics. According to Bouwmeester et al. [2], the characteristic thickness can be expressed theoretically as: Lc - D D ks - D* k8 ks (tel = 1) (2.2) where ks is the surface exchange coefficient, D, is the self-diffusion coefficient of oxygen anions with valence charge zo (= -2) and D* is the tracer diffusion coefficient. Ds is assumed to be identical to D* if the correlation between D* and ks can be neglected. D* and k, values can be obtained from the 180160 isotope exchange, permeation and/or coulometric titration experiments [6,74]. For equation 2.2 to be valid, a small oxygen partial pressure gradient is assumed across the membrane, and electronic conduction is assumed to be predominant. Bulk Transport The Wagner theory (eqn. 2.3) is widely used to depict bulk diffusion for a relatively thick membrane (i.e. L > Lc) [2]. Jo =-- RT 42 F 2 L 6 Jin' PO2 del + dn P0 2 "l"P5 (2.3) Cion Jo2 is the oxygen flux, R is the universal gas constant, T is the membrane temperature, F is the Faraday constant, and L is the membrane thickness. P and P" represent the high and low partial pressures of oxygen, respectively. o-e and o-in are the partial electronic and ionic conductivity, respectively. Equation 2.3 has been derived assuming: " The overall oxygen permeation rate is determined by the diffusion of oxygen ions through the membrane lattice or the transport of electronic charge carriers within the bulk oxide [2]; and " Two charged species exist in local equilibrium with a hypothetical neutral species within the bulk oxide. An example is oxygen ions and electrons in equilibrium with molecular oxygen [6]. Bulk diffusion is influenced by a minimum of three species. They are (a) oxygen vacancies, 41 (b) electrons, and (c) electron holes. Any of them could be the slowest moving species and hence have the strongest influence on the oxygen flux. A detailed analysis of these species is given by Surnaso et al [6]. Surface Exchange Surface exchange reactions on the membrane surface are two-fold. On the feed side of the membrane, molecular oxygen combines with available oxygen ion vacancies to give oxygen ions and electron holes thus: 1 -02 +V" 2 - (2.4) 0 0 +2h' On the sweep side, the surface reaction is the reverse of that observed on the feed side. The oxygen ions and electron holes combine to give molecular oxygen leaving oxygen ion vacancies in the membrane thus : 0 0 + 2h' 1 -02+ V** 2 (2.5) The transport of oxygen ions within the bulk phase can be improved by a reduction in membrane thickness or an increase in either conductivities (ionic or electronic) of the membrane. If this occurs, the limiting step becomes surface exchange kinetics and a new expression is required for the oxygen permeation flux al. [70], [75,76]. According to Bouwmeester et the Wagner equation is not suitable for a surface exchange-limited process of oxygen permeation through the membrane. Instead, the expression becomes [2]: 1 teltiono-total o 1 + (2Le/L) tel and tio 42 F 2 APo2( L are the electronic and ionic transfer numbers, respectively. o-total is the total conductivity. In deriving equation 2.6, identical surface exchange rates were assumed for both the feed and sweep sides. The membrane characteristic thickness (Lc) is expressed as [2]: L = ' where je RT teltionTtotal j.o Jex 42F2 (2.7) [mol.cm- 2 s-1 ] is the balanced surface exchange rate when oxygen potential gradients are absent, and its value can be obtained from 10-6 0 isotopic exchange data. The mechanism of surface exchange is expected to be a combination of sub steps in sequence which may have the capacity to individually limit the overall oxygen permeation rate. On the feed side, these sub steps include adsorption of oxygen molecules to the membrane surface, and the transfer of charges between adsorbed species and other species within the membrane surface. The reverse of these sub steps will occur on the sweep side [68]. A detailed analysis of possible mechanisms to incorporate the surface reaction kinetics into the bulk-diffusion mechanism is given by Du et al. [71]. 43 Overall Oxygen Flux Equations Several research groups have developed single expressions for the overall oxygen permeation flux across perovskite membranes [67,77,78]. These expressions incorporate all possible limiting processes during oxygen permeation. According to Liu et al. [5], the flux relations for planar/disk-shaped and tubular perovskites are given in Equations 2.8 and 2.9 respectively. krDv (P Jo 2LKf Dv R kkr [ (B (Ro - Ri) -k ) 2 )0.5 5(Pb + Rm/ 2 0.5] _ ( - (P2) )05] P2 (p2)0.51 Rin -D (Pb) (2.8) -P2 (Pb - 2 ( P2 (0bP )05 + D k,(P0P 02 =2 - 2 0 5 05 (2.9) + Rm/Ro -D (Ps Dv is the diffusion coefficient of oxygen vacancies, K, and Kf are the experimentally obtained reverse and forward surface reaction rate constants, respectively. R0 , Ri, and Rm are the outer, inner and log-mean radii, respectively. Rm = (Ro - Rin)/ln(Ro - Rin). According to Surnaso et al. [6], the assumptions that make Equations 2.8 and 2.9 valid, are: 1. Since the ionic conductivity of a perovskite is much less than its electronic conductivity, the flux of oxygen vacancies is assumed to govern the rate of oxygen permeation. 2. Electron holes move fast enough to make the electric field gradient non-steady (i.e. electronic conductivity is governed by the electron holes). 44 3. The gas phase exhibits the behavior of an ideal gas. 4. For a tubular membrane, radial diffusion is negligible and the oxygen vacancy diffusion coefficient is assumed constant. 5. Equation 2.4 represents the feed side surface exchange reaction while equation 2.5 represents the sweep side surface exchange reactions. Also, the forward and reverse surface exchange rate constants, kf and k, (given in equations eq:transport-equationdisk,eq:transport-equation-tube), are assumed to be equal and applicable to the surface reactions. 6. The membrane electron hole concentration is assumed constant for both surfaces. The other species (O, 02, V"*) are analyzed with the law of mass action. 2.2 Reactive ITM Applications ITM reactors have been used for different reactive purposes. Membrane technology has advanced beyond separation of oxygen only. Researchers have reported the coupling of oxygen separation and reactive process in a single step in ITM reactors. The advantage of the coupled process is that the mechanisms for oxygen separation and fuel reaction can be studied in a dependent-form analysis [79]. In this section, the main reactive applications for ITM reactors, which use methane as fuel, are discussed. A comparison of these applications as well as the challenges associated to them, are also presented. In general, the three reactive ITM mechanisms (for methane conversion) differ as follows: Oxidative Coupling of Methane: Takes place ON the ITM surface. CH 4 reacts with oxygen ions to form CH 3 , which couples to produce C2 H and C2 H4 . The ITM is used to limit availability of 02 and prevent oxidation to CO or CO . 2 Syngas Production: Partial oxidation (fuel rich) of CH 4 with oxygen (though usually with the aid of a catalyst) to produce CO and H2 . The reaction can also occur on the reaction surface with molecular oxygen. Oxymel Combustion: Oxymel combustion requires sufficient 02 for complete oxidation to CO 2 and H2 0. Catalysts will improve conversion efficiency at lower temperatures. 2.2.1 Oxidative Coupling of Methane Oxidative coupling of methane is the formation of C2 products by reacting CH 4 and oxygen ions. Although C2 H6 and C2 H4 are the main C2 products, C24 is mostly desired [8]. All permeated lattice oxygen is consumed on the membrane surface and gaseous oxygen is absent in the reaction side as shown in Figures 2-9 and 2-10. CH4 Air Side C2 , COx -- 2V0+02(g) 2o"+4k Membrane V;I oI Reaction side Z Mt aTy-st Reacdon Side layer -* H 2 0+2VG+2CHc oxygensiFe 0+2k'+2CH4 (s) )(3) C2HI(9) +h U)+ 02(g) Cil 4(8) TI 02 1120 + C 2 H 4 (g) + V; -3C Figure 2-9: Surface reaction using a catalyst layer [7] 2 H1 (j) + 00'+ k Figure 2-10: Reactions during oxidative coupling of methane [8] Three trends are common with most oxidative coupling of methane (OCM) experiments: 46 1. There is preference for tubular membranes with OCM-enhancing catalyst (usually Labased) packed in the tube side [38]. 2. Use of a reaction inhibitor to prevent total oxidation catalytic activity of the membrane. A good example is presented by Lu etal. [37] in which BaCeo. 6 SmO. 4 0 3 _6 catalyst was used to prevent the total oxidation catalytic activity of a SrFeCoo. 5 0 3 6 membrane reactor, by coating the inside of the membrane tube with the catalyst. 3. The use of membranes with good catalytic properties for OCM at lower oxygen partial pressure. Good results for C2 yield (17%) were obtained with B1 . 5Y 0 3 SmO. 2 0 3 _s (BYS) and BaCeo.8 Gdo.203 2.2.2 membranes [37, 45, 46]. Syngas Production Of all the potential applications for mixed conducting ceramic membranes, partial oxidation of methane (POM) to syngas (CO + H2 ) is thought to be one of the most commercially important applications [80]. The partial oxidation of methane is a promising alternative for syngas generation because it is a mildly exothermic reaction and could produce syngas with the H2 /CO ratio of 2:1, which is a preferable feed stock for methanol synthesis or the FischerTropsch reaction [49]. Combining oxygen separation and partial oxidation of methane into a single step is therefore important for cost reduction of syngas generation and the processes which use syngas [81]. Production of syngas in an ITM reactor involves fuel rich reactions which favor the formation of CO and H2 [8]. One side of the membrane is exposed to an oxidizing atmosphere 47 (usually air) and the other side is exposed to a reducing gas mixture (CH 4 , CO, H2, C0 2, H2 0 and so on). Therefore, it is crucial that the membrane materials should have a stable lattice structure under a wide range of oxygen partial pressure and high resistance to the highly reducing atmosphere besides possessing the desired oxygen permeability [8]. As with OCM experiments, three trends are common with most syngas experiments: 1. Preference for hollow fiber membranes [35]. A major advantage of this configuration is its relative large surface area to volume ratio. 2. Use of reforming catalyst on sweep side and reduction catalyst on feed side (Fig 2-11). Perovskite films and Ni-based catalysts (such as Ni/Al 2 0 3 ) are commonly used for reforming [35]. Platinum is commonly used as oxygen reduction catalyst. 3. Use of membranes with high oxygen permeability to further reduce the ITM reactor volume and cost [8]. Syngas (CO + H2) Oxygendepleted air CH4 (steam) Ai Air Reforrning catalyst 02 reduction catalyst ITM Figure 2-11: Production of syngas using an ITM reactor [9] Oxymel Combustion 2.2.3 Methane oxymel combustion involves reacting methane with permeated oxygen to ideally produce carbon dioxide, water vapor and heat thus: CH 4 + 202 -9 CO 2 + 2H2 0 (g) + heat (2.10) The reaction is exothermic and the products can easily be separated by condensing the water vapor thereby capturing carbon dioxide for reuse in the fuel stream or sequestration. Oxymel combustion requires sufficient oxygen for complete oxidation which in turn requires the appropriate CH 4 /0 2 molar ratio (i.e. 1/2). Catalysts may also be used to improve conversion efficiency at lower temperatures, as lower temperatures improve membrane stability. The use of ITMs for methane oxymel combustion reactions is a relatively new concept. et al. [15] Tan carried out oxymel combustion in a hollow fiber catalytic LaO.6 Sro. 4 Co 0 .2 Feo. 8 03-6 (LSCF6428) membrane reactor. Tan et al. prepared a hollow fiber membrane reactor (HFMR) using four LSCF6428 hollow fiber membranes (1.1mm I.D, x 27cm) with the insides packed with catalyst made of the granular powder (0.6mm dia.) of the membrane material. The membranes had an asymmetric structure (i.e. a thin separating dense layer integrated with porous layers on one or both sides). A comparison is also made with the same configuration that includes the effects of platinum catalyst on the air feed side (outside of the membranes). During the experiments, the methane feed concentration was kept at a constant 9.75% vol. (CH 4 with balance He). 2.2.4 Comparison of ITM methane applications It is important to point to the major differences in the ITM investigations discussed earlier in order to further understand their purpose and uses. Table 2.1 outlines the major differences. Table 2.1: Comparison between ITM investigations in oxidative coupling of methane (OCM), syngas production and oxymel combustion. Criteria Reaction location OCM Preferred membrane geometry Catalyst for fuel conversion P02 (Partial pressure of oxygen on the Tubular Syngas Mostly in the gas phase Hollow fiber Hollow fiber La-based (e.g La/MgO) Lower than syngas case Ni-based (e.g Ni/Al 20 3 ) Higher than Oxymel and OCM cases Granular form of membrane in use Lower than syngas case 800 - 950 0 C > 850 0 C > 850 0 C Membrane surface Oxymel Gas phase fuel side) Currently used operating temperatures CH 4 inlet concentration Desired products Challenges Lower than syngas Higher than Oxymel case and OCM cases C 2 HA, C2 H6 CO, H2 e Increase of temper- e Developing ITMs ature beyond 950'C with high 02 flux and leads to COx prod- long term phase staucts bility Lower than syngas case C0 2 , H2 0 e Development of ITMs suitable for, and stable at power plant operating temperatures 0 Higher methane * Chemical deposiconversion rate usu- tion and membrane ally leads to lower C2 instability have been selectivity reported ('-'-i 1400-C) * Developing catalysts that aid complete combustion of methane in the gas lys phase It can be seen that unlike OCM, syngas production and oxymel combustion requires reactions that occur mostly in the gas phase. Oxymel combustion is similar to syngas production in terms of membrane geometry preferred and best operating temperatures. It should also be noted that the methane feed concentration and partial pressure of oxygen on the fuel side are usually higher in syngas production than in oxymel combustion and OCM. 2.2.5 Reactive ITM challenges The future of reactive ITMs looks promising based on early results and projections. However, there are challenges that remain which must be overcome. According to Dyer et al. [82] the challenges for ITM applications in general include: e the development of systems that can incorporate the ITM technology; e the fabrication of mechanically, chemically, and thermally stable ITMs; e the development of mechanically, chemically, and thermally stable sealants with excellent sealing performance; e the development of economically-viable technology for ITM fabrication; and e long-term performance and dependability. The challenges facing reactive ITMs on the fundamental level are explained below: Oxidative Coupling: A major challenge for OCM reactions is the C2 yield. Although OCM experiments have been ongoing for over two decades and efforts have been put into developing novel catalysts, C2 yields are mostly reported to be < 25% [8]. Although an increase in temperature would favor the formation of more ethylene (C2H4 ), too high temperature (>950'C) would introduce more oxygen into the gas phase leading to the formation of COx products [50]. 51 Syngas Production: Very high temperatures are harmful to reforming catalysts [35]. Also, researchers have reported membrane contamination instability during syngas reactions [8]. The latter is mainly due to exposure of the membrane to an oxidizing atmosphere (usually air) on the feed side a reducing gas mixture (CO + H2 ) on the permeate side. Oxymel Combustion Little work has been done on oxymel combustion. Therefore the challenges that lie ahead would likely be more than those currently faced during OCM or syngas reactions. One major challenge that would be encountered during oxymel combustion is the selectivity of CO 2. Since carbon capture is one of the main goals of oxymel combustion, the presence of C2 products and/or CO (however little) in the product stream would limit the overall success of any oxymel reaction since carbon capture would no longer be possible by simply condensing the water vapor after reaction. Another challenge is the unsuitability of currently used ITMs for the temperatures (~ 1400'C) 2.3 used in natural gas power plant combustors. Methodology for ITM Reactor Characterization The use of ITMs within reactors for oxygen separation and reaction with gaseous fuels is a developing trend. Recently, researchers have developed well defined methods for ITM reactor characterization. The methods for sealing and leak detection are well documented while experimental methodologies which govern the reactor purging, temperature, inlet gas flow rates, fuel dilution and post-operation membrane analysis are well understood. The methodology for analysis involves the calculation of oxygen flux, methane conversion, and the selectivities or yields of reaction products. This section serves to provide an insight into the methods in current use for ITM reactor characterization. 2.3.1 Sealing and Leak Detection The sealing of the ITM reactor is an important step in preparation for experiments mainly because the oxygen flux expected is very low (in the order of 0.1pmol.cm--2.-1), and as such, any unexpected gas species within the reactor would invalidate the results obtained. Sealing is however a major challenge in ITM reactors partly because of the general lack of reliable seals at the very high temperatures required for oxygen flux, and the differing thermal expansion coefficients between membranes and supports. However, some progress has been made in optimization of reactor sealing and reduction of leaks across the seals. ITM reactors reported in literature tend to be small-scale laboratory setups. As a result, the areas that require sealing - reactor connecting parts and the membrane - are few and small in size. The common trend is to apply custom-made gold, silver, or glass-based rings to these areas and achieve sealing at a high temperature, close to the melting point of the sealing material [83]. In most cases, an almost perfect sealing of the reactor is possible. The leak across the membrane seal is generally quantified by calculating the leakage of air or oxygen based on the concentration of nitrogen or other trace gas detected on the sweep side at temperatures varying from room temperature to 900'C [50]. The value of nitrogen leak (as part of air) into the reactor sweep side is usually obtained using a gas analysis equipment such as a gas chromatograph. The values of the oxygen (as part of air) and total air leaks are obtained 53 using equations similar to 2.11 and 2.12 respectively as shown below: ( 0.21 V0 2 ,leak = . Vair,leak = 7 9 VN 2,leak 1 VN 0.79 (- (.2 2,leak (.2 where it has been assumed here that the molar fractions of oxygen and nitrogen in air are 0.21 and 0.79 respectively. V0 2 ,leak, VN2 ,leak, and Vair,leak stand for the leak flow rates of oxygen, nitrogen, and air respectively. The quantified oxygen leak is considered when determining the oxygen flux (see section 2.3.3). Zhu et al. [84-87] reports the use of gold rings to seal a dual-phase membrane reactor at 1040*C for POM experiments, and silver rings to seal membranes at 960'C for oxygen permeation experiments. Schlehuber et al. [88] reports that when gold rings were used for sealing a disk-shaped Lao.5 8 Sro. 4 Co 0 .2 Feo.8 03-3 (LSCF) membrane in a reactor which operated at 800'C for 3000 hours, the leakage level of oxygen was less than 5% of the oxygen permeation flux. Tan et al. [15] reports the use of high temperature water-based glass-ceramic sealant to achieve sealing in a hollow fiber membrane reactor (HFMR) which uses a Lao.6 Sro. 4 Coo. 2 Feo.sO (LSCF6428) membrane and achieves oxymel combustion. This results in a leak which is 0.4% of the oxygen permeation flux. Wang et al. [38] reports <0.2% leak using a ceramic glass binder, while Zhang et al. [89] reports that leakage was below the detection limit when glass rings - produced from glass with a high melting point - were used. However, glass seals have been reported to become highly reactive at high temperatures leading to failure of the seal after several hours of operation [90]. Glass seals could also diffuse into membranes at 36 high temperatures thereby contaminating parts of the membrane and causing errors in the calculation of oxygen flux and other parameters (e.g. fuel conversion) [29,91,92]. 2.3.2 Experimental Methodology Some experimental methods are common during ITM reactor operation. They include the techniques developed for membrane heating rate and temperature range, flow rate of inlet gases, fuel dilution, and post-operation membrane analysis. Temperature Temperature control is intrinsic to the operation of any experimental ITM setup. It is generally achieved by placing the ITM setup in a temperature controlled furnace (generally tubular). The rate of increase of membrane temperature is expected to be limited by the thermal expansion of the various reactor components, as well as the volume of reactor material to be heated. If considerable time is required for heat equilibration, the heating of the membrane and the reactor will need to be slow to avoid high temperature differences within the setup. Most heating rates reported in literature tend to be relatively slow - the most commonly reported heating rates are between 1 - 50Cmin-1 [15,48,93]. The typically preferred temperature range is 700 - 900'C. Inlet Gases Due to the low oxygen flux derived in ITM experiments (in the order of 0.l 1pmol-cm-2.s 1 ), the flow rate of inlet gases are relatively low. They are controlled using mass flow controllers 55 upstream of the reactor inlets. In some cases, the inlet conditions are kept constant, but in most reactive cases, the feed and sweep flow rates are varied to study their effects on oxygen flux and the thermochemical processes. The flow Reynolds number (Re) range is typically 1 to 25. Fuel The fuel (i.e. methane) used in ITM reactors is usually diluted with an inert gas (usually Helium or Argon) to control the temperature and stoichiometry during reaction. This is usually done by mixing the fuel and diluent streams after setting the needed flow rates through their respective mass flow controllers. Reported fuel inlet concentrations vary from 0% to 100%. In most cases where the fuel concentration is kept constant during experiments [15,46,50], the value is usually <10%. Some cases however exist in which the sweep side gas is always 100% fuel for reactive experiments [85, 94]. The most common trend is to vary the fuel inlet concentration in order to study its effects on oxygen flux and the macroscopic thermochemical processes. Post-operation Membrane Analysis In most cases, membranes are analyzed for phase, structure and surface changes after their use. The techniques commonly used are X-ray Diffraction (XRD), Scanning Electron Microscopy (SEM), and Energy-Dispersive X-ray spectroscopy (EDX). These techniques have helped identify membranes which are stable in reactive environments and also to study the effects of surface changes on the type of reactions which occur. 56 XRD yields the atomic/crystalline structure of materials and is based on the elastic scattering of X-rays from the electron clouds of the individual atoms in the membrane. SEM is used study the membrane surface topography and conductivities. The composition of the membrane is determined using EDX. 2.3.3 Methods for Analysis In order to obtain the values of oxygen flux and other macroscopic thermochemical processes, researchers tend to use gas analyzers and a mass balance analysis based on the type of experiment. Some researchers use an oxygen sensor/analyzer to determine the oxygen concentration in gas streams [50]. However, the most commonly used gas analyzer is the Gas Chromatograph (GC). GCs use a capillary column, a detector, and carrier gases to separate and quantify the components of a gas mixture. The concentrations obtained from oxygen meters and GCs can be used to calculate the actual values of oxygen flux and other processes as needed. The most commonly values are oxygen flux, conversion of reactants, and the selectivity or yield of products: Oxygen Flux The oxygen flux (Jo 2 ) calculations are usually based on the type of experiment/reaction - separation, oxidative coupling, syngas, and oxymel. For oxygen permeation experiments in which leakage through the membrane seal is considered [88], equation 2.13 is used. Here, the oxygen leak has been incorporated into the flux equation. 57 In experiments for oxidative coupling of methane and syngas production [48, 50], equations similar to 2.14 are used. In this case, the oxygen flux is simply evaluated based on the change in oxygen concentration in the air stream and its increase in the sweep stream. For oxymel experiments, the equations would be similar to other reactive cases except for the fact that the removal of water vapor before entry into the GC complicates the analysis. Therefore the equation for oxygen permeation in oxymel reactions is estimated based on the stoichiometric coefficients of the reactions to the carbon oxides and C2 products. An example is shown in equation 2.15 [15]. The unit of the oxygen flux in the equations below is cm 3 (STP)/min. ( J o2 J0 2 Jo2 where Xo 0.21 0 = 0. 7 9 7feed,in X .79 1 = VSout (Xo 2 ,S,out+ XN 2 ,s,out leakage, - sout 2 (2.13) Vsweepin N2 Sout (1 - (XO2 ,8,OUt/100) - (XN 2 ,s,out100)) (2.14) X0 2 ,a,out - X0 2 ,a,out 2 xco 2 ,S,oUt + 1.5xco,out + XC 2 H 4 ,S,OUt + 0.5xc 2 H6 ,S,OUt) (2.15) is the measured molar fraction of nitrogen on the reactor sweep side indicating ,S,out is the measured molar fraction of oxygen on the reactor sweep side, X0 2 ,a,out is the measured molar fraction of oxygen on the reactor air side, and Vsweep,out is the flow rate of the product gas stream. Conversion of Methane and Oxygen The conversion of fuel is one parameter of interest for an ITM reactor since a high conversion rate would usually indicate that the oxygen flux obtained is near ideal for a given fuel flow rate. It can be easily calculated as shown in equation 2.16 [15]. The conversion rate of oxygen is however a bit more complicated. It is calculated based on the oxygen concentrations in the air and sweep exit streams as shown in equation 2.17 XCH4= (2.16) XswepouCHsou X 100% Vweep,inxCH4,s,inX0 - 2 ,aout) X02 = 100 [50] 1 - sweepinX Veein 0. 21 Vfeca,in 21 0.79Vfeed,inXO 2 ,a,out/ (1- 2 sout (1 - 202o) (2.17) where XCH 4 XCH 4 ,s,in represents the methane conversion, X0 2 represents the oxygen conversion, and represents the mole fraction of methane in the sweep gas inlet stream. Product Yields and Selectivities Depending on the type of desired reaction, the selectivity and yield of a certain specie are commonly used parameters for characterizing the ITM reactor performance. In the case of 0CM, C 2H4 is generally considered while for syngas production and oxymel combustion, CO and CO 2 are considered respectively. The analysis is based on the assumption that methane is the fuel feedstock. Equations 2.18 and 2.19 [50] represent the yield (Yi) and selectivity (Si) of any carbon59 containing specie i. Y = 100 x i"s'o't (2.18) XCH 4 ,s,in S, = 100 x * = 100 X XCH 4 i ,sout XCH 4 ,s,in - (2.19) XCH 4 ,s,out where ni is the number of carbon atoms in the molecule of a carbon-containing product i. 2.4 Analysis of ITM Experimental Investigations ITM reactor technology has developed impressively since the mid-1980s. As discussed in sections 2.1 and 2.2, there are many different membranes, membrane geometries and structures, and applications. As a result of these developments, common trends are observed with ITM reactor operating methods and processes. The differences in membrane types, operating conditions, reactor geometries, inlet gases, and applications present a set of challenges when an analysis into the overall governing processes is to be conducted. Some of these challenges include large differences in oxygen permeation flux (up to 3 orders of magnitude difference) and a lack of obvious trends for fuel conversion and species selectivity for reactive analysis. This section serves to provide an in depth analysis of ITM reactor governing processes. The analysis is focused on normalization of oxygen flux using temperature and an effective activation energy, to consider the effects of temperature and sweep side mass transfer, as well 60 as fuel conversion and CO selectivity in the case of reactive sweep gases. Thirty experimental investigations have been studied in the development of the analysis in this section as shown in table 2.2. An absolute (total) pressure of latm is assumed in both the feed and sweep sides for all the investigations. Table 2.2: Experimental investigations from literature considered for analysis No. Author Year Experiment ITM Name (geome (geometry) 1998 0CM LSCF-8264 (disk) 2 1 Zeng et al. [40] 2 Lu et al. [37] 2000 0CM 3 Zeng et al. [45] 2000 0CM SFC (tubular) BY25 (disk) Feed Gas ITM Sealant V.G & Silver 50% 02 + 50% N2 Helium Air Helium Pyrex glass Bi2 03 -based Inert Se Sweep Gas 50% 02 + 50% N2 Helium T (reactive and non-reactive) T b c BSCF (disk) 2001 0CM BSCF (disk) Gold 2001 0CM BYS (disk) Glass & BYS powder 2002 Syngas BCFZO (disk) BCFZO (disk) Gold [95] 2001 5 Shao et al. [96] 6 Zeng et al. [46] Tong et al. [97] SCFO a 7 ONV UNK3 Syngas Dong et al. 4 c 50% 02 b Yaremchemko et al. [98] 10 Va,in = 200ml V0 'in (reactive) (reactive) Air Helium T (reactive) Helium T (reactive and non-reactive) Helium T (reactive and non-reactive) Vs,in (reactive) + 50% N2 T (disk) Air 2003 Syngas LSCG (disk) UNK Air Helium 2004 Syngas LCFC (disk) Gold Air Argon (non-reactive) [99] SCFO (ITM) BCFZO (ITM) BCFZO (ITM) T (reactive) (reactive) VTi T (non-reactive) Diethelm et al. = 300ml & XCH 4,in zGH4,in disk ITM tubular ITM a T (non-reactive) b Va,in (non-reactive) T = 800"C c Va,j, (non-reactive) T = 850*C d Va,in (non-reactive) T = 875*C a (non-reactive) V, Vs,in (non-reactive) T T = 875'C f Wang et al. [100] 2004 Separation-only BSCF (tubular) Air Ceramic Glass Helium = 800*C T = 850"C g Vs,in (non-reactive) a T (non-reactive) b T (reactive) 1% CH4 c T (reactive) 3% CH4 at inlet 11 Ikeguchi et al. [48] 2005 Syngas SBCF (disk) Pyrex glass Air T (non-reactive) Vn b 12 at inlet XCH 4 ,on Argon a 13 V 0 ,i, T (reactive) Helium a b T (reactive) Air a 8 (non-reactive) T (reactive and non-reactive) a 3 Other Varied Parameter Liu et al. [101] 2005 cV Separation-only BSCF (hollow fiber) Air Silver Helium 0 i Vsi, (non-reactive) T = 850*C (non-reactive) (non-reactive) T = 900*C T = 950*C a T (reactive and non-reactive) no catalyst b c CHnno T(ctiv (reactive) catalyst catalyst L LaSr/Ca catalyst Wang et al. [38] 2005 0CM BSCF (tubular) Ceramic glass Air or 0 2 /N2 mixture Continued on Next Page ... Helium o No. Author Year ITM Name Experiment ITM Sealant Feed Gas (geometry) Inert Varied Parameter a T (non-reactive) hb t e (non-reactive) HFMR(A), T =850*C 0 (non-reactive) HFMR(A), T = 900-C T f h ST Liu et al. 14 [771 2006 Separation-only BSCF Silver (hollow fiber) Air Helium 1 V,,j, (non-reactive) V,,,i, (non-reactive) (non-reactive) 1, a b c Tam et at. (5O] 2006 16 h Zhu et at. [86] 2006 0CM LSCF-6428 (hollow fiber) Syngas BCF1585 (disk) hV,,j Ceramic glass Gold Air Argon V0 , ,_ V0 i. Air Helium T HFMR(C) HFMR(C), T =850*C HFMR(C), T =900*C HFMR(C), T =950*C (reactive) (reactive) (reactive) rective HM() T = 900*C T (non-reactive) L 0.4mm b Vs,,.,(non-reactive) L 0.4mm Ti (reactiecie)FRB) T(o-ecie Vs,in (non-reactive) L .m 0* 0.8mm c d Zhamg et at. [69] 2007 a Syngas LCCF (disk) S a Glass Air Helium 0 ,i, ds)V V 0 ,i (reactive) (reactii~e) V V, 'in i (nonreactive) V0 ,i (reactive) d e f b a b9 V0 ,in =l0ml V0 ,, 20m1 T (reactive) c 18 diskB)T =95* (reactive) (reactive) T b Tam et at. [15] 2008 Oxymel Zu [86]1 20106 Syngas et al. a LSCF-6428 (hollow fiber) BF58 Ceramic glass Air Argon Vi,, (reactive) V,,., (reactive) (disk) Gold Air Helium V0 ,in (reactive) VnI(ecive T (reactive) CHn Vr,i, a Diethelm et at. 21 aa Kozhevnikov at at. [102] 22 b Wamg et at. 2009 Separation-only BSC(tubular) Glass Air Argon 845*C T =845 T Pt -4*C T catls 950"C = 2009 Syngas LSFTa1 (disk) Gold Air Argon 2009 Syngas LSF (ubula) Glass Air Argon Vs,,, 2009 Syngas 8CFZ (hollow fiher) Gold Air Helium T (reactive) T (reactive) (non-reactive) T(ecie dVContin(reactivex) [103] [49] T a b c T T =900*C5" HM() T =950*C T (non-reactive) Buysse et at. [56] 20 24 HFMR(B), T = 950*C ha a 16 = 850oC HFMR(B), T =9OO*C T (reactive) 15 17 HFMR(B) HFMR(B), T (non-reactive) k 1 HFMR(A), T =950*C (non-reactive) Vj, (non-reactive) V,,i. (non-reactive) V0,,j 0 (non-reactive) g HFMR*(A) 0, V,,(non-reactive) d 1 Other Sweep Gas Va,i Tianm et al. [105] 2010 Syngas SC-LSC (tubular) Glass Air UNK (reactive) (reactive) (reactive) (reactive m V,in ) (reactiven)Te=845* Tag=.90" tubula IT HFMr(C) No. Year Author Experiment ITM Name Inert Feed Gas ITM Sealant Varied Parameter T (non-reactive) a b Va, j (non-reactive) T = 750*C c V 0 ,i, (non-reactive) T = 800*C d Va,in (non-reactive) T = 850*C e Vain (non-reactive) T = 900*C f V,j (non-reactive) T = 950*C g h Vs,in (non-reactive) T = 750*C V, (non-reactive) T = 800*C V0 ,in (non-reactive) T = 850"C (non-reactive) (non-reactive) T = 900'C Wei at al. 25 [106] 2010 Separation-only BSCF (hollow fiber) Helium Air Glass V,i, V,,in k Zhu et al. 26 [85] 2010 Syngas SDCSSF (disk) Air Gold Helium T a (reactive) XCH T c 27 Gong et al. [107] 2011 Syngas LBFZ-0.2 (disk) Silver Helium Air 4 ,in (reactive) XCH d 4 ,in (reactive) L = 1mm L = mm L = 0.5mm L = 0.5mm 10% Ni/ty-A1 2 0 3 catalyst 10% Ni/-y-A1 2 0 3 catalyst a T b Vin c T (reactive) -y-A12 03 catalyst T (reactive) LSCF catalyst d e Kniep et al. [93] 2011 Syngas SCF (disk) Silver 20X 02 + bal. Argon (reactive) Vs,in Helium (reactive) XCH 29 Markov et al. [94] 2011 Syngas LSF (tubular) Glass Air Argon b 2011 Syngas BSCF (disk) Ceramic Glass Air Helium T Shen et at. [108] 4 ,in LSCF catalyst LSCF catalyst xCH 4 ,in a 30 T = 950*C T (reactive and non-reactive) b 28 Other Sweep Gas (geometry) T (reactive) (reactive and non-reactive) 'Used as sweep gas in non-reactive experiments and/or as fuel diluent in reactive experiments. 2 V.G = vacuum grease 3 UNK = unknown. ITM without laser ablation ITM with laser ablation 2.4.1 Temperature Analysis The effect of temperature increase on oxygen permeation flux is perhaps the most understood process in ITM reactor operation. Figure 2-12 shows the plots of Log (J0 2 ) vs 1000/T for reactive and non-reactive analysis. s seen from the plots, oxygen flux increases with temperature. It can also be seen that rr ost investigators consider membrane temperatures above 700'C. The experimental setups which use barium-containing membranes exhibit the highest oxygen fluxes. In order to carry out further analysis of flux dependence on reactor operating conditions, it is necessary to normalize the flux across all investigations. This is necessary because the fluxes reported vary up to 3 orders of magnitude, and because of the differences in each reactor. Since oxygen ion migration through perovskites is temperature activated [89], oxygen flux can be related to temperature via: AO x J0 2 = exp ( (2.20) where Ea is the activation energy and AO is a pre-exponential factor which accounts for parameters including the membrane thickness L, oxygen partial pressures (P 2 andP), surface reaction constants (kf and kr), and conductivities (owrn and -e). E is the slope of Log (J0 2 ) vs 1/T. Log (AO) is the y-intercept. For the purpose of flux analysis, normalized oxygen flux J Jt*> _ J2 Ao x exp 65 (E. (- 2 is obtained thus: (2.21) ----\-----1Or- 0 CO E . E *Wang -2 - 0 ------ O 0 -- T- Zeng1998 Lu2000 Zeng2000 Zeng2001 Tong2002 (a) Tong2002 (b) Diethelm2004 (a) 2004 (a) Ikeguchi 2005 (a) Liu2005 (a) Wang2005 (a) Liu2006 (a) Liu 2006 (e) Liu2006 (i) 9Zhang 2007 (a) - Zhang2007 (c) - Buysse2009 (a) - Diethelm2009 (a) Wei 2010 (a) Zhu 2010 Shen2011 (b) -5 -6 . 0.7 0.8 0.9 1 1000/T [K~1] 1.1 1.2 1.3 (a) 3 ----- Zenag 1998 2- 0-- .- - 0.Tan E .. J... Ikeguchi .... ) E -1 -2 ~. 2 - 0 -j - -3- 0.78 5I 0.8 I 0.82 ' I 0.84 0.86 A Luo 2010 (a) - Tian 2010 (a) Zhu2010 Gong2011 (a) Gong2011 (c) Kniep2011 (a) Kniep2011 (c) - Kniep 2011 (d) Shen2011 (a) -Shen 2011 (b) i I Dong 2001 (a) Dong 2001 (b) Shao 2001 Tong2002 (b) 2005 (b) ]keguchi 2005 (c) Wang 2005 (a) 2006 (a) Zhu2006 (a) Tan2008 (a) Tan2008 (b) Diethelm2009 (a) I 0.88 0.9 - 1 0.92 : I 0.94 ' I 0.96 a _ 0.98 1 1000/T [K ] (b) Figure 2-12: Reported effects of temperature on oxygen permeation flux for: (a) separationonly cases; and (b) reactive cases. Figure 2-13 shows the ratio of reactive to non-reactive pre-exponential factors for five reported investigations. The ratio is greater than 1 as expected (since the reactive flux is higher in all cases). In the case of Shen et al. 2011, the non-reactive flux reported barely increases with temperature while the reactive flux increases by a factor of 2 from 750 - 850*C. Furthermore, at 850*C, the reactive flux is an order of magnitude more than the non-reactive Shen 2011 (b) Zhu 2010 Wang 2005 (a) Tong 2002 (b) Zeng 1998 0 0.5 1 AOreactive / 1.5 2 2.5 3 Onon-reactive Figure 2-13: Ratio of reactive/non-reactive pre-exponential factors flux. This is most likely responsible for the large pre-exponential factor ratio reported by Shen et al. 2011. In figure 2-14, normalized oxygen flux is shown as a function of temperature for both reactive and non-reactive cases. As expected, the flux values are fairly close to 1 in most cases. This shows that the normalization method, albeit imperfect, is quite suitable for comparing investigations with different membranes. The exception occurs with Liu et al. 2006 (a) in which the flux showed a slow increase up to 700C and a rapid increase immediately after. This could be due to an order-disorder transition within the membrane oxygen vacancies as temperature is increased past a certain temperature [91,92,109,110] (700C in this case) which would cause a change in Ea and is not reflected in figure 2-12. It should be noted that, for the most part, many reactive ITM processes are yet to be fully understood. However, of all the reactive processes, the dependence of fuel conversion on temperature is the most understood. From figure 2-15, it can be seen that the fuel conversion increases with temperature for all reported cases. In all cases, the fuel inflow was kept at a constant rate, and oxygen permeation increases with temperature (as discussed 67 1.4 Zeng1998 Lu2000 Zeng 2000 Zeng2001 Tong 2002 (a) Tong2002 (b) Diethelm2004 (a) Wang 2004 (a) keguchi2005 (a) Lu 2005 (a) Wang 2005 (a) Liu 2006 (a) 3 1.3 --- 1 1.1 Liu2006(i) Zhang2007 (a) 0.9 2007 (c) -Zhang Buysse 2009 (a) Diethelm2009 (a) Wei 2010 (a) 0.8 -Zhu 2010 Shen 2011 (b) 550 600 650 700 750 800 850 900 950 1000 1050 Temperature, T [0 C] 1.15 -Zang 1998 Dong2001(a) Dong2001(b) Shao 2001 1.1 Tong 2002(b) -Ikeguchi * 2005(b) Ikeguchi 2005(c) " 1.05 * I 0.95 - - 0.9 Shen2011(a) 0.85 0.8 700 Wang205(a) Tan2006(a) Zhu2006(a) Tan2008(a) Tan2008(b) - Dietheln2009(a) SLuo 2010 (a) Tian2010(a) Zhu2010 Gong2011(a) Kniep2011(a) Knip 2011(c) Kniep2011(d) Shen2011(b) 750 800 850 900 Temperature, T [*C] 950 1000 1050 (b) Figure 2-14: Normalized fluxes as a function of temperature for: (a) separation-only cases; and (b) reactive cases. earlier). Therefore it can be assumed that the increase in fuel conversion is mainly due to the increase in permeated oxygen available for conversion of the fuel as temperature increases (and accelerates oxidation kinetics). 68 80 60 20 - c 0 700 750 800 850 900 Temperature, T [(C] 950 1000 Figure 2-15: Reported effects of Temperature on CH 4 conversion 2.4.2 Mass Transfer Analysis Mass transfer effects in ITM reactors has been studied by varying the inlet feed or sweep flow rates. Most of the analysis however focuses on the sweep side. The general consensus is that an increase in the feed flow (mostly air) which already has the higher oxygen partial pressure, will barely impact the local feed partial pressure at the membrane surface. pin 00ea Therefore, the chemical potential gradient and oxygen permeation flux across the membrane barely increases with increasing feed flow. The mass transfer analysis in this section studies the effects of Reynolds number and residence time of the sweep gas. For the purpose of proper analysis, only reactors with tubular or hollow fiber membrane configurations were analyzed since characteristic lengths and active volumes are difficult to estimate (based on published data) for disk JTM setups. The Reynolds numbers (Re) for the reported investigations are calculated via: Re- (2.22) where p is the gas density, V is the inflow rate, and p is the gas viscosity. De is a characteristic reactor length which represents the internal diameter for tubular and hollow fiber membrane reactors. The gas residence time is calculated via: =.(2.23) V where V is the active volume of the reactor: this represents the membrane volume in tubular or hollow fiber reactors. -- 2.5 Wang 2004 (e) Wang 2004 (f) - Wang 2004 (g) - Liu 2005 (b) - 2 -*g - Liu 2005 (c) - -.-.- - -7-y- Liu 2005 (d) 2006 (b) . . . .Liu . o~ 1.5 Liu 2006 (c) ----- Liu2006(d) Liu 2006 (f) Liu 2006 (g) Liu 2006 (h) -- V- Liu 2006 (j) E-- 1.- z -v- Liu 2006 (k) --- Liu 2006 (I) - 0.5 - 0 2 4 6 10 8 12 14 16 Buysse 2009 (b) Wei Wei Wei - Wei Wei 2010 (g) 2010 (h) 2010 (i) 2010 (j) 2010 (k) ReResweep H- Figure 2-16: Non-reactive oxygen flux dependence on Reynolds number Figure 2-16 shows the effect of sweep Reynolds number on oxygen permeation flux in non-reactive cases. Oxygen flux increases with Reynolds number but begins to level off gradually as Reynolds number increases beyond a value of 5. It can be assumed that the mass transfer resistance approaches its minimum value after Re ^ 5. 70 7 1.6 1.6 1.4- 1.4 7 1.2 -- Tan Tan -A Tan ,-A Tan -A--Tan 1.2 1 2006 (c) 2008 (f) 2008 (g) 2008 (h) 2008 (i) 0.80. 0 Z -4-Tan 2006 (c) 0.6 0.4 0.2 0 Tan 2008 (f) -A-Tan 2008 (g) -A-Tan 2008 (h) --A-Tan 2008 (i) 0 5 10 15 Rsweep 20 z 0 0.6 0.4 25 0.20 ___, _______........________ 1 2 Residence Time (sweep) [s] (a) 4 (b) Figure 2-17: Reactive oxygen flux dependence on: (a) Reynolds number; and (b) residence time. The sweep gas is a mixture of an inert gas (He or Ar) and CH4 . Figure 2-17a shows the effect of sweep Reynolds number on oxygen permeation flux in reactive cases. Oxygen flux shows a similar behavior to that observed in the non-reactive cases. The oxygen flux also begins to level off beyond Re a 10. In this case however, it is beneficial for the residence time to be higher than the non-reactive case. This is because the fuel will need to react with the permeated oxygen. As seen in figure 2-17, the residence times are generally higher than the non-reactive cases. 100 -4-Tan 2006 (c) Tan 2008 (f) -A Tan 2008 (g) 80 Tan 2008 (h) -- Tan 2008 (i) .2600 o4020 01 0 5 10 15 20 25 30 Resweep Figure 2-18: Dependence of CH 4 conversion on sweep mass flow rate. As with the case of temperature, fuel conversion shows a common trend as sweep mass flow is increased. An increase in the sweep mass flow rate decreases the methane conversion in all reported cases shown in figure 2-18. This is because the reactant (CH 4 ) residence time within the reactor becomes shorter with increasing mass flow and increasing fuel/0 hCH4 ,s,in/7h02 ,perm) 2 (i.e. ratio. Further increase in fuel inflow will increase the oxygen permeation, but is detrimental to the conversion of the fuel. 2.4.3 Reactive Analysis ITM reactor analysis under reactive conditions involves a number of factors which cannot be compared easily (between different setups) by simply studying the effects of temperature, mass flow, or fuel dilution. Amongst other things, the different fuel flow rates, fuel dilution percentages, catalysts, and temperatures present different reaction pathways, and mechanisms within the ITM reactors. However, a study of the effects of the fuel/0 2 ratio for all reactive cases of temperature, mass transfer, and fuel diluent variations reveals some trends for fuel conversion and the selectivity of CO (syngas experiments only). The performance of these reactors under reactive conditions can also be studied in comparison with equilibrium. An equilibrium state implies the reactant and product concentrations do not change over time. Equilibrium was calculated using Cantera (an open-source chemical kinetics software). For the purposes of all reactive analyses in this study, the fuel/0 2 ratio is the ratio of fuel inlet molar flow rate to the molar flow rate of permeated oxygen thus: 72 fuel _ nCH ,s,in (2.24) 4 n 0 2,perm 02 Zan1g 1990 2001(a) 099 9 2001( b) 100 U Dong 200 1 (d) 90 80 00o 2001 Tong 2002(0b) Tong 2002c (a) D1eft 2004 2000)0) Ik'g000i 2000(9 Wang 20050(2)l To,,20060(9) Tn 2006 ( T20200( Zhu 2000(2a)) Zhu 2000 (b) -kogo 40 60 Too 2000 (a) Ton 2008 (b) Kno 2000 () "Cd .0 CD 0 100 5a2 o0 Tan20080(0) a 2000 (f) 0+ Toan 2000 () CH : 40LTo 'r 30 20 21) 00) Looh 2010 )0 02 L.02010 (0) L4 00201 () Gong2011(0 A 20 9000g2011(d) KiP21 Knip 201 o Kniep2011)b 0 00 ' ' ' B' ShoO() I 5 10 15 Eqoh0lbiun 00'C CH 4 :0 2 [- Figure 2-19: Dependence of CH 4 conversion on fuel/0 2 ratio. Figure 2-19 shows the dependence of fuel conversion on fuel/0 2 ratio. Some of the investigated ITM reactors operate close to equilibrium. In some cases however, the methane conversion is far below equilibrium. One distinction that can be made is between disk and tubular/hollow fibers. It can be seen from the plot that most disk reactors tend to achieve equilibrium state of fuel conversion while tubular/hollow reactors do not. This could be attributed to the longer fuel residence time in disk reactors. For Tan et al. 2006 and 2008, it is possible that unknown reactor conditions contribute to the low fuel conversion. In the case of Kniep et al. 2011 (c), blank '}-Al 2 0 3 was used in place of high performing catalysts and this contributed to poor methane conversion. In some other cases (such as Zeng and Wang et al.), OCM experiments were carried out leading to low fuel conversion (since there is insufficient oxygen in the gas phase). 73 0 0 0 V 100 90 V 80_ Diethelm 2004 (a) Ikeguchi 2005 (b) Ikeguchi 2005 (d) U Zhu 2006 (a) 6 Zhu 2006 (b) U Kozhevnikov 2009 (a) * Kozhevnikov 2009 (b) * Kniep 2011 (a) * Kniep 2011 (b) Kniep 2011 (c) Kniep 2011 (d) 2 Kniep 2011 (e) * Kniep 2011 (f) X Markov 2011 0 Equilibrium 850 C 10 -- - Equilibrium 9000 C X 6050 40 30 20** 10 -' 0 0.1 (a) (b) (d) (b) (c) * - 70- o Dong 2001 Dong 2001 Dong 2001 Tong 2002 Tong 2002 1 CH4 : 02 Figure 2-20: Dependence of CO selectivity on fuel/0 2 ratio. Unlike the methane conversion, reported CO selectivities are mostly close to equilibrium as shown in figure 2-20. There are some slight overestimates (e.g. Kozhevnikov et al. 2009) but it can be seen that high CO selectivity is generally achieved with increase in the fuel/0 2 ratio between 1 - 6. Kniep et al. 2011 (c-f) are the only major exceptions: in these cases, LSCF and -y-A1 0 2 3 catalysts were used instead of the high performing syngas catalyst (Ni/y-A1 2 0 3 ) used in Kniep et al. 2011 (a-b). 2.5 Conclusions Significant research has been done in the experimental characterization of ITM reactors for oxygen separation and fuel combustion over the last three decades. The investigation of ITM reactors for methane conversion purposes has become a common trend due, in part, to the benefits of combining oxygen separation and methane combustion in one unit. 74 The three major groups of methane conversion experiments are oxidative coupling of methane (formation of C2 products), production of syngas (CO + H 2 ), and oxymel combustion (desired products are CO 2 , H2 0 and heat). A common experimental trend involves studying the effects of important operating parameters - temperature, methane and air flow rates, and methane feed concentration - on the oxygen flux, selectivities and yields of products, and methane conversion; though the understanding of the underlying processes and how these parameters impact them is still somewhat open. These operating parameters are the most important factors in determining the best operating conditions in ITM reactors. A number of challenges with the use of ITM reactors have been identified. They include, but are not limited to, low C2 yield in OCM reactions, damage of catalysts and membranes at very high temperatures and reducing atmospheres (syngas), and difficulty in producing pure CO 2 stream (oxymel). An analysis of reported ITM experimental investigations provides an insight into the governing processes for ITM reactor operation. As expected, oxygen permeation flux is favored by temperature increase. The reactor sweep inflow rate also improves the flux, but its potential for increasing the flux further is limited as minimum mass transfer resistance approached. For reactive mass transfer analysis, increase of fuel flow rate is beneficial to oxygen permeation but detrimental to fuel conversion as the fuel/0 2 increases. An analysis of the fuel conversion and CO selectivity shows that some reported investigations operate close to equilibrium based on the fuel/0 2 ratio. However, factors such as poor catalysts or lack of sufficient oxygen in the gas phase could limit the reactor performance. Currently used ITM reactor setups tend to be small-scale laboratory setups. The analysis carried out is mostly 'global' and focused on the inlet and exit temperatures, concentrations, 75 pressures and flow rates. An in depth study into the effects of the oxygen-fuel ratio, the specific reactions which occur, and a spatial resolution of the reaction zone will provide further insight into ITM reactor operation. Chapter 3 Experimental Approach The development of the experimental ITM set up was done with the aim of fulfilling the following goals: 1. Design, construction and assembly of a bi-directional stagnation flow reactor for study into the response of oxygen flux and other processes (e.g. fuel conversion) to global reactor operating parameters (e.g. flow rates, temperatures, fuel dilution, etc). 2. Spatial resolution of species within the reactor with the aid of optical and temperature measurement techniques. Experiments were carried out in the Reacting Gas Dynamics laboratory, making use of an ITM reactor. The reactor was designed using commercial software application (SolidWorks). A MATLAB code was developed for the data acquisition system (National Instruments PCI 6229) which controls and monitors operating conditions of the reactor and records data from diagnostic equipment. MATLAB was also used for post-processing of data. 77 The experimental approach covered in this chapter provides an insight into the design of the ITM reactor and its process control, the membrane used, and the methodologies developed for experimental procedure and analysis. 3.1 Reactor Design and Installation The most important step in the study of ITM reactors is the proper design and installation of said reactor. The design of the ITM reactor was done based on the configuration of Figure 3-1. The design ensures air flows in from the top while the sweep gas flows in through the bottom of the reactor. As a result, during reactive experiments, the reaction zone occurs at a stagnation region just below the membrane. After separation/reaction, oxygen-depleted air flows through the top manifold while the reaction/separation products flow through the bottom manifold. The reactor configuration is therefore one-dimensional (along the stagnation streamline), thereby facilitating a fundamental study of processes involved during separation and/or reaction. Importantly, the configuration aids cross validation with in-house numerical 1-D modeling work done along the stagnation streamline. 3.1.1 Reactor The ITM reactor was designed for laboratory-scale oxygen separation and non-premixed combustion experiments. It operates at atmospheric pressure and is housed in an enclosure (discussed in section 3.1.2) to, among other things, minimize heat losses during experiments. The reactor is made entirely of Inconel 601 - a specialty metal mainly composed of Nickel 78 ITM ITM C02, H20 Optical access Reaction Zone streamline CH4, C02 Figure 3-1: Basic schematic of the ITM reactor (58-63%), Chromium (21-25%), and Iron (remainder). One reason for using Inconel 601 is its stability in reactive reducing environments and at the temperatures needed for ITM reactions (currently >700'C). The stability of inconel gives it an advantage over stainless steel which is stable up to 600-700'C and deteriorates/oxidizes under reactive conditions. Inconel is also relatively easy to maintain and machine (unlike ceramics). Another reason for designing the entire reactor body with inconel (bolts, nuts, panels, e.t.c) is to prevent uneven thermal expansion that would lead to cracks, leaks and rust within the reactor. The reactor (without its enclosure) is shown in Figure 3-2. The maximum dimensions of the reactor are 49" x 5" x 50 (L x B x H). Air flows in through the top manifold and fuel/diluent flows in through the bottom manifold while four exhaust manifolds provide the flow path for the exhaust gases. Mass flow controllers (±1% accuracy) are used upstream of the reactor inlets to provide a maximum of 2000sccm, 500sccm, and 50sccm for air, CO 2 79 inlet manifold Pyrometer view access SamplingAir connection (4 sampling points) Sensor connection manifold Sensorexhaust Sensor connection Exhaust manifold for sepa ration/reaction products Cartridge heater Optical access Fuel/diluent Inlet manifold Figure 3-2: The ITM Reactor (diluent) and CH 4 (fuel) respectively. There are 10 connection points on the exhaust manifolds for sensors - pressure sensors, thermocouples, oxygen sensor, and gas chromatograph. The exhaust manifolds have 4 points each. The sampling lines used at these points are of 1/16" tubing (except the oxygen sensor which is 23mm in diameter). There are four bores for cartridge heaters which provide fine control of temperature. A sight tube which is 430 from horizontal allows access to the pyrometer (installed outside the enclosure) to read the membrane temperature at the center of the membrane. Figure 3-3 shows the front section of the reactor and its main part (before exhaust). The inlet manifolds are fitted with flow straighteners to allow for uniformity of flow characteristics along the stagnation streamline. Five sampling connections provide access for 80 Flow Straightener Cartridge Membrane Heater 1.25"P Sampling connection (4 sampling points) 0.88"1 0.5" (For adjustment of sweep side gap height) Thermocouples Figure 3-3: Reactor front cross-section thermocouples and Gas Chromatograph (discussed in section 3.3.3) sampling lines. The topside sampling point provides access to analyze the feed composition and temperature, while the four sampling points on the bottom provide access to the reaction/separation gas species. Sampling can be done normal to the membrane (between membrane and exhaust manifolds) on both sides thereby measuring the gas temperatures before separation or reaction. Sampling can also be done anywhere between the flow straighteners and the exhaust manifolds on the sweep side, thereby aiding spatial resolution and numerical analysis. Each sampling point can take up to four sampling lines (1/16" tubing) . Four other small inlets allow for thermocouples to read the temperature of the center plate at four different points. The gap for the feed side within the reactor is about 1.25" while that for sweep side is about 0.88" (adjustable to 1.38"). The adjustability of the permeate side allows for studies 81 into the effects of gap height on the type of reaction and species formed. The 0.88" gap also enhances mass transfer (for the low inlet flow rates used). A future adjustment can be made to adapt the reactor for parallel flow experiments by sealing off the current inlets and using one exhaust manifold on each side as inlet for feed and sweep gases. 3.1.2 Reactor Enclosure Enclosure heater terminal Pyrometer access Exhaust manifold Optical access Inlet for cartridge - Enclosure insulation heaters, sensors, and sampling lines 32" Figure 3-4: The ITM Reactor within its insulation As earlier stated, the main function of the reactor enclosure is to prevent heat losses from the reactor. It is made of Gemcolite@ No-Smoke@ FG23-103 material and is stable up to 1260 0C. Figure 3-4 shows the reactor within its insulation. The outer dimensions of the enclosure 82 are approximately 32" x 25" x 39" (L x B x H) and it is 4" thick. Just like the reactor, there are two optical accesses made of quartz glass on opposite sides of the enclosure and an access duct for the pyrometer. The cartridge heaters, sensors and sampling lines go through four holes drilled into the insulation. The enclosure also houses the reactor support bricks and six enclosure heaters. The enclosure heaters provide radiant heating within the enclosure to the reactor body and membrane (see section 3.3.4). 3.1.3 Reactor Sealing The sealing of the ITM reactor was one of the most challenging tasks during installation and commissioning. This is mainly due to gases escaping through metal-metal and membrane-metal joints and partly due to the manifold flanges deforming during bolt tightening. A breakdown of the different sealants used is important and pressure/leak tests which help verify the integrity/performance of the sealants is presented below. Sealants For the ITM reactor, different sealants were used for the membrane, manifolds (connections between manifolds and reactor main body), and other joints. Membrane: Alumina felt (supplied by fuelcellmaterials.com) which is stable up to 1650'C, is used to seal the membrane to the reactor center plate. Figure 3-5a shows the sealant on the topside of the membrane. It is applied top and bottom of the membrane and installed in the space shown. The sealing of the membrane reduces the available surface area on both sides from 3.4" x 3.4" to 3" x 3". It should be noted that membrane 83 Membrane installation Point within Reactor (a) Membrane Inlet manifold Gasket Reactor Caulked Joint (ceramic adhesive) (c) Caulked Joint (b) Gasket Figure 3-5: The different sealants used for the ITM reactor sealing is the most crucial sealing within the ITM reactor because any non-quantified leaks across the sides of the membrane from feed to sweep side would invalidate the values of oxygen flux and other variables obtained during experiments. Manifolds: Gaskets (Flexitallic@ Thermiculite® 867) are used for the connections between manifolds and reactor main body. The gasket material is composed of a stainless steel core (support) and mineral coating (sealant). It provides stiff sealing for the reactor joints. As seen in Figure 3-5b, the gasket is designed with as little surface area as possible in order to maximize seal preload. 84 Other Joints: A water-soluble ceramic adhesive (Cotronics® Durabond® 952) is used to 'caulk' the ungasketed joints, as seen in Figure 3-5c. It has similar thermal expansivity to Inconel 601 and is therefore a near-ideal sealant at high temperatures. It takes about 2 days to fully caulk the reactor: (1) the preparation and application of the adhesive takes about 10 hours, (2) the caulk is allowed to set for 24 hours at room temperature, (3) the caulk becomes stronger when allowed to set for 2 hours at 1000C and an additional 2 hours at 200'C. The order of overall reactor sealing is thus: " Using alumina felt, seal the membrane to the center plate; " Apply ceramic adhesive to the ungasketed joints. This is called caulking; " Allow the caulk set at room temperature for about 24 hours; " Assemble the reactor parts with gaskets used at the manifold joints; and " Heat up the reactor and allow the caulk strengthen by keeping the temperature constant at 100'C and at 200'C for 2 hours each. Sealant performance tests Before proceeding to carry out experiments using the ITM reactor, it was important to test the integrity of the sealants used. A basic pressure test along with the use of leak testing fluid would help indicate the leak areas along the body of a reactor. However, further leak tests are required to quantify the actual leaks within the reactor and through the membrane seal. Purging >2000sccm CO2 Infiltration test 1 02 sensor Air? >2000sccm CO2 250sccm CO2 Membrane seal test Infiltration test 2 4" 500sccm Air Air? 500sccm CO2 500sccm CO2 Figure 3-6: Sealant performance tests Sealing (at 250C) After Sealing (at 250 C) After Sealing (at 210 C) 0.9 -Before 0.8- - -- 0.7 - - 0 E 0.6- - - - 0.5 S0.40.3 0.2 0.1 -- 0 .... Infiltration Test 1 -..... Infiltration Test 2 Membrane Seal Test Figure 3-7: Sealant performance results: quantified in ymol.s-' of oxygen (from the air which leaks into the reactor sweep side) The tests were carried out at three different stages of reactor sealing development: (A) before application of the sealants (measurements at room temperature, 25'C); (B) after application of the sealants (measurements at room temperature, 25*C); and (C) after application of sealants (measurements at 210C). The reactor was purged with pure CO 2 before each test. The oxygen sensor was used to monitor and measure the increase of oxygen 86 0.9 Membrane Leak Test nfiltration Test 2 0.8 8 - 0.7 N E 0.6k E 0.5 ME 5-- 0 0.4 a) (U 0.3 (21--e 0.2 0.1 0 Before Sealing (250 C) After Sealing (250 C) After Sealing (21 00 C) 0 Figure 3-8: Comparison of total leak at 500sccm air and 500sccm C0 2, to the expected oxygen permeation flux (assumed to be in the order of 1pmol.cm- 2 .s 1) concentrations on the reactor sweep side caused by any leak of 02 (in air) into the sweep stream. Therefore, the air that has leaked into the reactor sweep side from the surrounding environment (infiltration tests 1 & 2), or from the feed side (membrane seal test), can be detected. All measurements were carried out after the value of oxygen concentration detected is equilibrated. The methodology for quantifying the leak (under non-permeating conditions, at temperatures below 600'C) is explained in appendix B. Figure 3-6 shows the test configurations and the order in which the tests are carried out: Infiltration test 1: The purpose of this test is to detect air leakage into the reactor at low reactor flow rates, by quantifying the concentration of oxygen. It is carried out after both sides of the reactor are completely purged with CO 2 . Thereafter, the sweep side 87 is set to 250sccm CO 2 flow. For this test and infiltration test 2, the feed side is filled with CO 2 (no flow on the feed side). Infiltration test 2: This test serves the same purpose as the first infiltration test, but is carried out with a higher flow rate of CO 2 on the sweep side. Both sides of the reactor are completely purged with CO 2 , then the sweep side is set to 500sccm CO 2 flow. Membrane seal test: This test is used to quantify the leakage rate of oxygen across the membrane seal and is carried out immediately after infiltration test 2. The flow rate on the sweep side is maintained, while the feed side is set to 500sccm air flow. After the new increased value of oxygen concentration has equilibrated, the leak across the membrane seal is quantified by the difference between this new concentration and that obtained during infiltration test 2. As seen from the results in figure 3-7), the leak into the reactor is greatly reduced by application of the sealants [NB: the membrane was always sealed using alumina felt during all tests, hence the relatively constant value of membrane seal leak]. From infiltration test 2, we see that the leak into the reactor is improved by a factor of 4 after sealing. The leak into the reactor is however less predictable at lower flow rates (- 250sccm) as seen in infiltration test 1. It can also be seen that the membrane seal leak is a factor of 5 less than the reactor leak. Figure 3-8 shows the total leak into the reactor sweep side at inlet flow rates of Vair,a,in 500sccm and Vco2,,in = 500sccm. The leak flux has been divided by the membrane surface area to enable comparison with expected oxygen permeation flux. The total leak into the reactor sweep side (infiltration + membrane seal leak) after sealing, is about 0.23% of the 88 expected oxygen permeation rate across the membrane (1pumol-cm--2-1). 3.2 Membrane The membrane is the most important part of the technology explored in this work. Without it, this would only be an experimental combustion set-up. The membrane used for any reactive ITM investigation must be suitable for long term use in reactive environments. What follows is a brief overview of the membrane details and the pre-operation analysis work done on the membrane. 3.2.1 Membrane Details The membrane used for this work is a perovskite mixed ionic-electronic conductor (MIEC) supplied by Ceramatec Incorporated. Its chemical formula is Laa. 9 CaO.1 FeO 3-6 (LCF)and it was chosen for its balance between suitable oxygen flux and stability in reacting environments. According the U.S patent 6,492,290 BI [111], membranes with the chemical formula: (LnxCai-x)yFeO 3 -5 where Ln is La or a mixture of Lanthanides comprising La, and 1.0 > x > 0.5 1.1 > y > 1.0 have been shown to provide acceptable oxygen flux while being stable for syngas production. Membrane (LaO. 9CaO FeO 34 ) Sealing gasket placed here Optical Access Fuel/diluent Inlet manifold Figure 3-9: Reactor top cross-section Figure 3-9 shows the bottom part of the reactor exposing the membrane. The membrane sits about 1/8" above the optical accesses: there are two optical accesses, made of quartz glass, on opposite sides of the reactor to facilitate the monitoring, analysis and imagery of the reaction zone. The membrane surface area used for oxygen flux calculations (section 3.5) is 9in 2 or approximately 58cm 2 . There are three membrane thicknesses used: 2.2, 1.3 and 0.89mm. 3.2.2 Pre-operation Membrane Analysis Due to the importance of surface exchange reactions on the membrane surface as well as membrane stability under reactive conditions, it was necessary to analyze the membrane pre and post-operation using X-ray Diffraction (XRD), Energy-Dispersive X-ray spectroscopy (EDX), and Scanning Electron Microscopy (SEM) techniques. The analyses were carried 90 FeK LaL LaL CaK .710 1.40 2.18 2.83 LaL 4.20 3.58 4.98 5.1 6.30 keV Figure 3-10: Pre-operation EDX graph P =peak P C, 0 P p C P P a) P P 10 20 30 40 50 60 70 2 theta 80 90 Figure 3-11: Pre-operation XRD graph out at the MIT Electrochemical Energy Laboratory'. EDX provides the composition and chemical characteristics of the membrane as shown in figure 3-10. The EDX technique is based on the interactions between electromagnetic radiation and the membrane, and the analysis of X-rays emitted by the membrane in response 'The analyses were carried out mainly by Lei Wang. 91 Acc.V Spot Magn 15.0 kV 3.0 Det WD 10000x SE -2 Pm 10.0 Figure 3-12: Pre-operation SEM image to being hit with charged particles. XRD (fig. 3-11) yields the atomic/crystal structure of the membrane and is based on the elastic scattering of X-rays from the electron clouds of the individual atoms in the membrane. There are 7 major peaks observed, which help to identify the orthorhombic crystal structure of the LCF membrane. SEM provides an indication of the membrane surface topography and conductivities. The technique involves using a high-energy beam of electrons to interact with the membrane's atoms. The 2pm scaled SEM image for the pre-operation membrane analysis is shown in figure 3-12. 3.3 Reactor Process Control and Instrumentation This section provides an overview of the reactor plumbing, instrumentation, gas chromatography, heating, and safety. The process control and instrumentation (PC&I) for the 92 ABBREVIATIONS: 3P = Three-Phase B= Circuit Breaker Cart. = Cartridge DAQ = Data Acquisition Encl. = Enclosure M = Mass flow controller MP = Mass flow controller purge P = Vacuum Pump P-0 2 = 10% 02, Balance N2 F = Filter GC = Gas Chromatograph S = Control Switch SCR = Silicon-Controlled Rectifier K = Key Lock Sol = Solenoid R = Relay RO = Rotameter T = Temperature Controller VAC - Alternating Voltage SECTIONS (see line colors) (1) Plumbing System (2) Instrumentation (3) Gas Chromatography (4) Heating System (5) Safety Control Figure 3-13: Reactor Process Control and Instrumentation (PC&I) reactor was designed to facilitate the optimization of flow and heating processes, sensor positioning, as well as to aid the safety the reactor as a whole. The major sections of the reactor PC&I setup (all discussed in this section) are shown in figure 3-13. Each section can be identified by the color of the lines in the figure, and most Figure 3-14: Measurement locations on the ITM reactor sections are interconnected. At the center of control is the Data Acquisition System (DAQ) which is connected to all sensors, mass flow controllers (MFC), solenoids, cartridge heaters, the oxygen sensor, and the pyrometer. A computer using an in-house developed Matlab data acquisition GUI controls the DAQ. The major process control and instrumentation components of the ITM reactor set up are listed in table 3.1 Figure 3-14 shows a schematic of the major reactor processes and measurements which aid the development of analysis equations (section 3.5) and the characterization of the reactor. All temperature measurements (except for the membrane) are done using thermocouples. 94 Table 3.1: Process Control Equipment and Instrumentation Equipment/ Instrument Application Measurement(s) Manufacturer and Model Mass Flow Controllers Pyrometer Inlet flow rate control Vfeed,in, VCH 4 ,s,in, Sierra Smart Trak 2 Vco 2 ,s,in T Impac IGA300 Thermocouples Gas and reactor temperature Tfeed,in, Tfeed,out, Tsweep,in, Tsweepout, measurements Tenc Pressure sensors Gauge Pressure mea- Pfeed,out, Psweep,owt Oxygen sensor surement Oxygen concentration X0 Data Acquisition System Cartridge heaters Instruments and cartridge heater control Reactor heating National Instruments Enclosure heaters Reactor heating Temperature controllers Silicon-controlled rectifier (SCR) Gas Chromatograph Relay and SCR control Enclosure heater control Gas concentrations I SQUARED R ELEMENT SEU Starbar Omega CN77R352-C2 Spang Analog C-Series Agilent [t-GC 490 Quad Membrane temperature measurement 2 ,a,out, Xo 2 ,S,o 0 t XH 2 , X02, N 2 , iXCH 4 , The temperatures of the inlet gases just before the membrane (Tfeed,out Omega PX309 Bosch LSU4.9 Dalton Watt-Flex CO,, I C2 Hz the exits from the reactor Omega Type-G (Tfeed,in & Tsweep,in) and at & Tsweep,out) are measured. Also, the temperatures at different points in the reactor center plate (T 1 _4) and inside the enclosure (Tec) are measured to control the radiant and enclosure heaters respectively (section 3.3.4). The temperature of the membrane (T) at its center is measured by the pyrometer. Pfeed,out and Psweep,out are the pressures measured at the reactor exits. The oxygen sensor provides a realtime measurement of the equilibrium oxygen concentration in the sweep side (x0 2 ,8 ,out) while the gas chromatograph (CC) provides the concentrations of the various separation/reaction products in the sweep side exit. The mass flow controllers (MFC) control the inlet flow rates of gases 3.3.1 (7rifeed,in & Tmfsweep,in). Plumbing The reactor plumbing system is shown in figure 3-15. It incorporates multiple flow configurations for: " the desired experimental flow processes, " purging of air out of the reactor (using C0 2 ) before experiments, " purging of the reactor and its enclosure (using CO 2 ) in the event of an emergency, and " calibrating the oxygen sensor. This is done regularly in-between experimental runs. The procedures for flowing gases into the ITM reactor are detailed in appendix C. The mass flow controllers are placed upstream of the reactor inlets to control the inlet flow rates, and provide a maximum flow rate of 2000sccm, 500sccm, and 50sccm for Air, CO 2 and CH 4 respectively. The CH 4 solenoid ensures that methane flow can be stopped manually and through the DAQ while the enclosure solenoid provides a safety mechanism for quick purging of the reactor enclosure with CO 2 in the event of an emergency (see section 3.3.5). 96 Figure 3-15: Reactor Plumbing Schematic 3.3.2 Instrumentation The reactor instrumentation is composed of pressure sensors, thermocouples, the pyrometer, and the oxygen sensor. The pressure sensors indicate the absolute pressure on each side and can be used to detect membrane breakage. The other three instruments provide important data for reactor characterization. Figure 3-16 shows the locations where instruments can be placed in the reactor. -_I-1 _ Figure 3-16: The major reactor instrumentation Thermocouples Thermocouples are used for two purposes within the reactor, (1) to measure the temperature of the reactor center-plate, and (2) for measuring the gas temperatures within the reactor and at the gas exhaust. As seen in figure 3-17, four thermocouples (Omega type-G) are placed within the reactor center-plate, adjacent to the cartridge heaters. These thermocouples provide the temperatures (T1 ) that enable the monitoring of the reactor temperature. thermocouple connection for feed inlet therm ocoupie for sweep inlet membrane position cartridge heater cente -plate thermocouples Figure 3-17: Reactor thermocouples Figure 3-17 also shows a thermocouple placed to read the temperature of the inlet sweep stream (Tsweep,in) just before the membrane. This is also done on the feed side of the reactor. Both measurements aid in the calculation of the change in temperature and heat transferred to the gases before separation or reaction (for air or fuel/diluent respectively). temperatures at the exits (Tfeed,out The gas & Tsweep,out) are also measured using thermocouples. 98 It should be noted that thermocouples can also be placed in a variety of places within the reactor gas streams to enable detailed analysis of localized effects with the specific goal of providing validation measurements to the numerical work done on the ITM reactor in other investigations. Pyrometer The membrane temperature is an important part of the operation of any ITM reactor. As discussed in detail in section 2.2, variations in the membrane temperature influence oxygen permeation rate, reaction thermochemistry and stability of reactive ITMs. Distance from membrane to pyrometer mount location = 18" Angle of pyrometer Pyrometer View access view = 43' Membrane Figure 3-18: Membrane temperature measurement configuration The membrane temperature for the reactor in this work was set by the radiant and 99 enclosure heaters and was measured using a Pyrometer (INFRATHERM pyrometer IGA 300) which has a measurement range of 300 - 1300'C and sends a current output (converted to voltage) to the data acquisition system. Its accuracy is within 0.8% of the measured temperature + 1C, i.e. i(0.008*Tmeas. + 1C). The pyrometer is located outside the enclosure to prevent it from overheating. It is mounted at an angle of 430 from horizontal and about 18" away from the membrane as seen from the side section of the ITM reactor in figure 3-18. There are three sapphire glasses between the pyrometer and the membrane (two within the enclosure and one on Sapphire was chosen because of its spectral the pyrometer view access on the reactor). (transmission) properties. 15 0.9 1100 1000 900 CU E 10 C CLID800 E0.8- ~ E 700 08 - C E 600 a) 500- -0- Emissivity -G-Current 400 1 2 4 5 Output Voltage [V] 3 6 7 0.7 400 500 600 700 800 0 900 5 1000 Membrane Temperature [ C] (b) (a) Figure 3-19: Pyrometer calibration data: (a) membrane temperature vs pyrometer voltage output; (b) membrane emissivity and pyrometer current output vs membrane temperature = (62.5 * I + 323) /T) The pyrometer was calibrated using a furnace and adopting a geometric configuration similar to its installation configuration within the reactor and using the same membrane type (Lao.gCao.1FeO 3 -6) used for the experiments. The calibration was carried out using 100 increments of 50'C from 500 - 1000'C. As seen from the calibration data in figure 3-19a, the membrane temperature is linearly related to the voltage output of the pyrometer. This calibration data was incorporated into the Matlab code developed for the ITM reactor thereby enabling the control/monitoring of the membrane temperature through the data acquisition system. Figure 3-19b shows the relationship between temperature, pyrometer current and emissivity of the membrane. The emissivity (F) of the membrane drops from 0.884 to 0.768 as the membrane temper- ature is increased from 500 - 1000'C. The membrane emissivities were obtained, based on the manufacturer's recommendation thus: S T )4 (62.5*1+323)4 (3.1) (3.2) where I is the pyrometer current reading in mA, T is the membrane temperature in Kelvin, and T is the temperature of a black body (in Kelvin) which produces the same pyrometer current reading as the membrane. Oxygen Sensor The oxygen sensor is used to provide the realtime value of oxygen concentration in the reactor sweep side exit. It is therefore invaluable during reactor leak testing and oxygen permeation experiments. The oxygen sensor used is an automotive type Bosch LSU4.9 universal exhaust gas oxygen sensor (UEGO). The sensor can measure oxygen concentration within the range of 0 - 0.21 bar in exhaust 101 Exit duct 12 1 Oxygen sensor location #023mm 41' Entry duct $ 1/4" 2 , , 2.5 3 3.5 Voltage Output [V] 4 4 Figure 3-20: Oxygen sensor calibration: (a) calibration duct; (b) calibration data gases and it operates up to a gas temperature of 1030'C. Its accuracy is within 0.24% of the measured oxygen concentration. The sensor is calibrated using a duct with a similar cross-section as the reactor exhaust manifold (see Fig 3-20a). The method is to flow: 1. pure CO 2 (i.e. 0% 02), 2. nitrogen and oxygen mixture (10% 02), and 3. air (21% 02) through the duct and then record the equivalent values of voltage output from the oxygen sensor. The 0-10% range (shown in figure 3-20b) is the range of oxygen concentration expected within the reactor sweep side. As with the case of the pyrometer, the calibration data was incorporated into the Matlab code developed for the ITM reactor as well as the data acquisition system. 102 3.3.3 Gas Chromatography The measurement of species concentrations is intrinsic to the overall goal of ITM reactor characterization. These concentrations help determine the oxygen flux across the membrane, fuel and oxygen conversion, and the selectivities/yields of products. The detection of a certain species is also a general indication of the occurrence of a type of reaction. A gas chromatograph is used to carry out the concentration measurements. The Gas Chromatograph (GC) used is an Agilent fp-GC 490 Quad. It is used to characterize the oxygen flux (02) as well as reaction products (02, CH 4 , COX, C2 Hz, H2 ,...) by reading the concentrations of the gases in the exhaust streams. The GC requires the use of Ar and He carrier gases and is equipped with a thermal conductivity detector (TCD). The detector responds to the difference in thermal conductivity between the carrier gas and the sample components. The minimum detection limit for the GC is 10-20ppm and the sample time is ~ 30 seconds. It uses three columns: " Column 1 (molesieve 10m, Ar carrier gas): detects H2 " Column 2 (molesieve 10m, He carrier gas): detects 02, N2 , Ar, CO & CH 4 * Column 3 (Pora Plot Q 10m, He carrier gas): detects other gases (CH 4 , C0 2 , C2 Hx) Figure 3-21 shows the GC sampling schematic. In order to analyze samples within the reactor, it uses sampling tubes (1/16" tubing) which can be inserted through any of the sampling points explained in section 3.1.1. Therefore, it is possible to analyze the concentration of gases in the feed and sweep side exhausts. The sample gas is filtered before entry into the 103 He Ar Sample HF H20 to Vent - FR caJ Gas to Vent GC -I Membrane Filter 1/16" tubing V Figure 3-21: Reactor Gas Chromatography GC to remove any water (liquid or vapor). Data from the GC is read and recorded using Agilent EZChrom software. 3.3.4 Heating The heating of the ITM reactor is important to facilitate separation and reaction since these processes occur only at high membrane temperatures (currently >700'C). The reactor is heated up using two sets of heaters - cartridge heaters and enclosure heaters. The cartridge heaters ensure fine control of the reactor and membrane temperatures while the enclosure heaters provide additional uniform heating capacity. Figure 3-22 shows the major components of both heating systems. 104 - - S Key Lock - - - Silicon- Distribution - SCR Controlled Block -c- Rectifier Master R Switch Relay SRCircuit Breaker LI Tmpo .Circuit Tm'Breaker Cont120VAC 120VAC 4 Tem p. Cont. 120VAC 208VAC (3P) Figure 3-22: Reactor Heating Systems Enclosure Heaters The enclosure heaters, supplied by I Squared R Element Company, consist of six silicon carbide heating bars connected to three-phase power (two heating bars per phase). They are installed within the reactor enclosure to provide uniform heating to the reactor. As seen in figure 3-23, the heaters have a cold section supported by the reactor enclosure material, and a hot section fully exposed within the reactor enclosure. The enclosure heaters alone provide a total output of 15kW. The control system for the enclosure heaters, as seen in figure 3-22 consists of the heating bars, a silicon-controlled rectifier (SCR), and a temperature controller. The temperature controller provides a set point to the SCR using variable current control. The enclosure thermocouple (located in the top inside of the reactor enclosure) provides the enclosure temperature reading to the temperature controller. The SCR, which uses three-phase power, 105 Enclosure Heater terminal Enclosure Heater (hot section) Enclosure Heater (cold section) Enclosure Figure 3-23: The Enclosure Heaters is used to control the heaters. Cartridge Heaters The cartridge heaters consist of four Inconel Watt-Flex@ heating bars. They are supplied with single-phase power (120V) and are installed within the reactor center plate as shown earlier in figure 3-17. The heaters provide a total output of about 2kW, and are operated using on/off control, with the set point and process values given by the DAQ software. With the enclosure heaters installed, the role of the cartridge heaters is to provide additional heating and fine control of membrane temperature (due to their proximity to the membrane). The control system for the cartridge heaters as shown in figure 3-22 consists of the heating bars, a solid state relay and a temperature controller which receives the set point via the 106 DAQ GUI. The process value is the average plate temperature as calculated by the DAQ. 300 -&E 280- Cartridge Heaters Only (SP:2500 C) Cartridge Heaters (SP:2000 C) + Enclosure Heaters (SP:250oC 260o 240 - 220- -. . .-. -. - ..-.- U) 0. E 200 180160 140 . 0 ' . 5 10 ' 15 ' ' 20 ' ' ' ' I I 25 30 I I 35 Position [cm] Figure 3-24: Heating tests with the cartridge and enclosure heaters (SP = Set Point). N.B: The green vertical lines indicate the cartridge heater locations Figure 3-24 shows the results of heating tests carried out using the cartridge heaters and enclosure heaters. The results are shown as a function of position from the first heater (0cm) to the last (32.5cm). The reactor center-plate temperatures are provided by thermocouples as shown in figure 3-17. When the cartridge and enclosure heaters are combined, the entire reactor can be heated almost uniformly at the desired temperature. However, the cartridge heaters take longer to equilibrate on their own, as seen from the difference in temperatures between the cartridge heater and center plate locations during the cartridge heater test. 3.3.5 Safety Provisions In order to ensure safe operation, it was important to design a safety system for the ITM reactor. Figure 3-25 shows the safety network designed for the ITM reactor. The main control points are: 107 Key lock: This is used to arm or disarm the heating system and fuel solenoid. Without arming the lock with a key, the heaters cannot be operated by using their control systems or the DAQ. In the same way, the methane solenoid cannot be operated by its switch or the DAQ system unless the lock is armed. Master switch: This is used to immediately switch off the heating systems and methane solenoid in an emergency during reactor operation. Enclosure purge switch: This enables the flow of CO 2 to the enclosure, thereby facilitating the cooling down of the enclosure area and diluting any methane present, and ultimately reducing the possibility of any reactions or fire occurring within the enclosure. Enclosure Purge Switch Figure 3-25: Reactor Safety Control 108 What follows is a brief discussion of the possible hazards that can be faced while operating the heating system or flowing methane in the ITM reactor, as well as steps that can be taken to prevent them. Heating Hazards The most likely hazard that can occur with the heating system is overheating. It is possible for the heaters to exceed the set point temperature provided through the temperature controllers. In this case, simply pressing the emergency switch would shut down the heaters and the heating systems. Both controllers also have an upper limit alarm to prevent further heating. Another hazard, although less likely, is the burning of unwanted objects (e.g. cleaning cloth) within the reactor or its enclosure. In this case, the safety procedures described under Methane flow Hazards below should be followed. Methane flow Hazards During reactive experiments, it is possible for the methane to leak into the reactor enclosure and react with the air within the enclosure leading to a fire. It is also possible for an unwanted reaction to occur within the reactor. The following steps will help eradicate or mitigate any methane hazard within the ITM reactor setup: 1. Press the master switch. This turns off the heaters and methane solenoid; 2. Turn on the enclosure purge switch. This purges the reactor enclosure with CO 2 (The 109 mixture of CO 2 and other gases is sent into a vent using a vacuum pump); and 3. Purge both sides of the reactor using the purge procedure described in appendix C 3.4 Experimental Methodology and Procedures This section provides a discussion of the experimental procedures and techniques developed for characterizing the ITM reactor. The type of combustion that occurs in the ITM reactor is non-premixed in nature. Nonpremixed combustion would be limited by the mixing of the reactant since the chemical reaction rates are much faster than the diffusion rates. A major advantage of non-premixed combustion is that the formation of a thin reaction front can be assumed, as the timescale of the chemical reaction (Tchem) becomes smaller than that of diffusion (Tdfgf). In the case of the ITM reactor, the mixing of reactants is limited by the oxygen flux across the membrane and this further increases the diffusion timescale. As a result, unwanted conditions such as pyrolysis of the fuel or low fuel conversion rates could easily occur if proper methods are not developed for controlling and measuring the membrane temperature, inlet gas flow rates, and inlet gas concentrations. Two types of experiments were carried out using the ITM reactor: Permeation-only experiments: In this case, air (21% 02) flows through the feed side inlet while pure CO 2 flows through the sweep side inlet. Oxygen permeates across the membrane from the feed side to the sweep side due to the difference in chemical potential (or partial pressure of oxygen) across the membrane. Therefore, the feed side 110 exhaust manifolds would transport oxygen-depleted air while the sweep side exhaust manifolds would transport a mixture of O2 and CO2Reactive experiments: In this case, air (21% 02) flows through the feed side inlet while a mixture of carbon dioxide and methane flows through the sweep side inlet. Oxygen permeates across the membrane from the air side to the sweep side as with the permeation-only case. Ideally, methane reacts with the permeated oxygen on the sweep side to form carbon dioxide and water vapor according to equation 2.10. However, the products of combustion may also contain carbon monoxide, hydrogen, unconverted CH 4 , excess 02, and/or other hydrocarbons (HC). Therefore, the air side exhaust manifolds would transport oxygen-depleted air while the sweep side exhaust manifolds would transport the products of combustion. As discussed earlier, the performance of an ITM reactor for methane combustion is usually based on the yield or selectivity of a carbon-containing product. Similarly, for the work done in this study, the primary variables of interest or performance indices are: 1. Oxygen flux (J02) across the membrane; 2. Carbon dioxide yield (YC0 2 ), for reactive experiments only; and 3. Exhaust gas composition. The methods for deriving of these variables are outlined in section 3.5 (along with the methods for deriving selectivities and yields of carbon-containing species (C0 2 , CO, C2 ), and the conversion of methane). The major focus of the work presented in this study, is the 111 effects of temperature, inlet gas flow rates, and methane inlet concentration on the primary variables of interest. 3.4.1 Overview of experimental procedures All experiments and measurements are conducted at atmospheric pressure within and outside the reactor. Before every measurement is conducted, the reactor is purged on both sides with about 2000sccm CO 2 for about 5 minutes to remove any air within the reactor. All measurements are conducted at the temperature, flow rates, and CH 4 inlet concentration desired after allowing the gas concentrations equilibrate within the reactor. All measurements are conducted at membrane temperature (T) between 700-800'C, sweep flow rate (Vsweep,in) between 200-500sccm, and methane flow rate (VH 4 ,,,in) between 0- 20sccm as shown in table 4.1. The feed side inlet gas for all experiments is dry air. The sweep side inlet gas is pure CO 2 for permeation-only experiments, and CO 2 + CH 4 for reactive experiments. The CO 2 acts as a diluent in the reactive case. Mixing occurs after the mass flow controllers, and the fuel/diluent ratio is based on the volume flow rates obtained from the mass flow controllers. It is important to note that when experiments are conducted to study the effects of one parameter on the primary variables of interest, other parameters are kept constant. Table 3.2: ITM reactor measurements Experiment T (0 C) varies Vfeed,in Vsweepin VCHs,in Permeation-only Controlled Parameter (range) T (700-800-C) (sccm) 500 (sccm) 500 - Reactive Sweep,in (200 - 500sccm) VCH4,s,inh (0 - 20sccm) 800 800 500 500 varies 500 112 (sccm) - varies 3.4.2 Temperature Control The heating system for the ITM reactor serves to provide a two-fold heating function. The enclosure heaters ensure uniform heating within the reactor and its enclosure, while the cartridge heaters enable fine control. The reactor is subject to a heating rate of 1.2'C min-' to enable uniform heating of the reactor mass and membrane, and to avoid uneven expansion which could lead to membrane breakage or reactor leakage. As heating is carried out within the entire reactor setup, the membrane temperature is monitored by the pyrometer. Two thermocouples installed just below and above the membrane measure the change in temperature of the inlet gases, thereby aiding the calculation of the convective heat transferred to the gases along the stagnation streamline, and thus aiding in-house numerical analysis. Measurements are carried out between 700-800'C, mainly because the oxygen flux obtained below 700'C is too low. Also, because of the slow heating rate (1.20C/min), it takes about 10 hours to heat the membrane (and reactor) from room temperature to 700'C. For experiments which study the effect of the membrane temperature on the primary variables of interest, measurements are conducted using 50'C increments from 700'C - 800'C for the membrane temperature. 3.4.3 Inlet Gas Control The inlet flow rates reported for current experimental ITM reactors are usually very low (Re = 1 - 50). As discussed earlier (section 2.3.2), the main reason for these low flow rates is the low oxygen flux expected across the membrane during experiments (in the order of 1 113 ymolcm-2.s-1). In addition, high flow rates on either side of the reactor could potentially lead to: " A high concentration of soot formed from methane pyrolysis, due to exposure of the methane to high temperatures and low concentration of the oxidant; " Increase in unreacted methane in the product stream; and " Breakage of the membrane, especially if the gas pressures on both sides are unequal. For permeation-only experiments, the sweep inlet flow rate was increased using 100sccm increments from 200sccm to 500scem. For reactive experiments, the effect of fuel dilution was studied by using 5sccm CH 4 increments from 10sccm to 20sccm (while keeping the CO 2 inflow rate constant at 500sccm). 3.5 Methodology for Analysis The methodology for analysis involves the development of a simple set of equations for the macroscopic thermochemical processes within the ITM reactor. All analyses rely heavily on the mole fractions of gases measured at the end of the reactor exhaust manifolds. The mole fractions are measured using a Gas Chromatograph (GC). For all calculations, it should be noted that all H2 0 (vapor or liquid) is condensed out of the product gas mixture, and the gas mixture is filtered before entry into the GC (see section 3.3.3) 114 3.5.1 Permeation-only Analysis The main variable to be measured during permeation-only experiments is the oxygen permeation rate across the membrane, and the nitrogen concentration (for leakage quantification). For oxygen flux calculations to be made, we need to analyze the flow of gases in and out of the reactor sweep side. The flow rates of gases into the reactor are given in standard cubic centimeters per minute (sccm). The standard conditions for these flow rates are: * standard temperature (Tt) = 273.15K " standard pressure (Pt) = 101325Pa (or latm) The mass flow rate (in kg-s-1) of sweep gas (pure CO 2 ) into the reactor is therefore: Thsweep,in MCo2 Mo x X Vs8weep,in RXPst Tn (33) where R is the universal gas constant and M is the molar mass. We make three considerations in this analysis: Firstly, we must take into account the leak into the reactor sweep side. We do this by measuring the nitrogen concentration (xN2 ,s,Out) on the sweep side exit stream. The method- ology for leak detection and quantification in the reactor has been discussed in section 3.1.3 and appendix B. The leak (X0 2 ,1eak) for this analysis is quantified using: 0.21 X0 2 ,leak 0.79 zN2,s,out giving the actual concentration of permeated oxygen 115 (Xo 2 ,perm) (3-4) detected in the sweep outlet X0 2 ,perm = X0 2 ,s,out - where X02,S,out X0 2 ,leak (3.5) is the total oxygen concentration detected in the sweep outlet. The average molar mass (M) of the gas mixture at sweep side exit is: M= (x x M) (3.6) Therefore, average molar mass (M) of the gas mixture at sweep side exit (permeation-only) is obtained from M = (Xo2,Sout x M0 2 ) + (xN2,S,Out where M is molar mass, and XCo 2 X MN2 ) + (XC0 2 ,s, 0 2 x MCo2) (3.7) ,S,out is the concentration of CO 2 detected at the sweep side exit. The mass fraction of any product i is obtained from M. A (3.8) (mfo2 ,perm), as detected in the sweep side is mfi'S'Out = xiS,02t x The mass fraction of oxygen permeated therefore Tmfo 2 ,perm =£02,perm X< M (3.9) Secondly, using the law of mass conservation, we assume the mass flow rate of the gas leaving the reactor sweep side (rhsweep out) is the related to the mass flow rate of the gas 116 entering the reactor sweep side msweepout (hsweep,in) thus: msweep, in -i XO2,8,OUt - X Mo 2 M -- (3.10) N2,s,out X MN 2 R Thirdly, we quantify the molar flow rate of permeated oxygen into the sweep side thus: 7O 2 ,perm mfo 2 ,perm X msweep (3.11) out With all the above considerations, the oxygen flux equation for permeation-only experiments can be formulated. The oxygen permeation rate across the membrane (in mol-cm-2. s-1) is given by: JO2 = AO 2,perm (3.12) A Or for direct analysis based on available data, VCO2,in JC 2 (XO 2 ,8 ,out XCO - 0.1XN 2 ,s'oUt) 2 ,8 ,oUt _St Tst RA (3.13) where A is the membrane surface area exposed to gases on one side. 3.5.2 Reactive Analysis For the analysis of reactive experimental results, the focus of the analysis is again on the sweep side. The analysis must consider the water vapor which is formed during reaction and 117 removed before entry into the GC. The variables to be measured are: (1) oxygen permeation rate; (2) methane conversion; and (3) the yields and selectivities of specific products (GO , 2 CO, H2 0, and H2 ). The overall reaction on the sweep side, if there is a leak and/or equilibrium is not achieved, can be given as: nO2 ,perm0 2 + nCH 4 ,s,inCH 4 + nCo2,s,inCO2 + h0 2 ,leak (3.727N 2 + O Reactants t Products h02,s,0ut0 2 + hCH4 ,s,outCH 4 + nco 2 ,s,OutCO 2 + +nCoS,,oUtCO where ni,s,in + nH02,,utH2 + nH 2 O,s,outH 2 0 nsoUtC represents the molar flow rate of reactant i and + (g) (314) nN2,s,outN2 ni,s,out represents the molar flow rate of product i. The 02 on the reactant side is a combination of permeated oxygen and the oxygen that has leaked into the reactor sweep side. The GC provides the measured molar fractions (Xi,s,meas) of the products in equation 3.14 (except H2 0) if they exist in the product stream. The values of Xi,s,meas will be higher than the actual mole fractions (xi,8 ,out) because H20 has been removed before entry into the GC. Using the same concept as in the permeation-only case (see equation 3.3), we obtain the 118 mass flow rates of CO 2 and CH 4 (both in kg-s-1) into the reactor as: MCO2 X Pt X Vco 2 ,s,in (3.15a) R x Tst mCH 4 ,s,in MC2X - Pst x VCH 4 ,s,in R x X VCo2,S,in R x (3.15b) Tst Similarly, the inlet molar flow rates are: Pst riCo 2 ,s,in = (3.16a) Tt x VCH 4 ,s,in R x Tt Ps nCH4,,,in - (3. 16b) The mole fraction of any product i (xT,,,aSt) formed in the reactor sweep side is obtained from: Xi,s,out = Xi,s,meas X where nH o,s,out and Z ni,s,,ut 2 >i nh~s,out - nH 2 O,s out (3.17) 1:hi'S Out are unknown. The average molar mass of products is obtained from equation 3.6 while the mass fraction of any product i is obtained from equation 3.8. The mass flow rate of products is where h02 ,s,in in,out =ms,in + nperm + rIleak mhst =ms,in + (no0 2 ,s,in x M 0 2 ) + (hN 2 ,s,in (3.18) X MN 2 ) (oxygen permeation + leakage) and nN2 ,s,in ( nitrogen leakage, same as (3.19) hNs,out are unknown. Therefore, apart from C0 2, we quantify the molar flow rate of any product 119 gas i (i,,,,ut, in mol-s 1 ) as: me, hi,s,out = mfi,,,out x ,out (3.20) Using these molar flow rates, we can obtain the total molar flow rate of oxygen gas (permeation and leakage) into the sweep side as: ho 2 ,s,in = hO 2 ,S,out 1. + 1 nco,sout +± ~-nH20,s,out 2 O~~ 2 OO2,sout + (3.21) Using the leak assumption explained in the permeation-only case, we can obtain the molar flow rate of permeated oxygen no2,perm (n102,perm) as: 0.21. -- o 2 ,s,in - 0.79N2,sout (3.22) With the equations derived so far (eqns 3.17 - 3.22), mass and mole balances are applied to equation 3.14 (with a Matlab code) until all unknowns are obtained using the 11 equations and 2 constraints below: 120 Oxygen Balance: 2 2fnO 2 ,S,out + 2 hCO2 ,S,0 Ut + nH2 O,s,out nO2 ,perm + 2 nCO2 ,s,in + 2 nO2 ,Ieak nCH4 ,s,in + C02,sin + hCO,s,out (323a) Carbon Balance: n CH4 ,sout + ftC02,sout + hCO's,out + nC,sout (323b) Hydrogen Balance: 4 ACH4 ,s,in 4 - iCH ,s,out 4 + 2 nH2 0,s,out + 2 nH ,s,out 2 (3.23c) Nitrogen Balance: 3 72 7 . nO2 ,leak = hN2 ,sout (3.23d) Mass continuity: mCH4 ,s,in o 2 ,S,outMO2 + hCH4 ,s,oUtMCH4 +nCO,S,outMCO + ?H 2 nO (3.727MN2 + M 0 2 ) = + hCO2 ,S,outMCO2 + hH2 O,s ,o utMH2 0 + mCO2,s,in + hO2,permMO2 + , 8,o&tMH2 2 ,1eak nC,s,outMC + nN ,S,OUt MN + 2 2 (3.23e) Molar flow rate of 02 gas at sweep outlet: nO ,s,out = 2 X0 2 ,s,meas X Ths,meas (3.23f) Molar flow rate of CH 4 gas at sweep outlet: nCH ,sout = 4 XCH 4 ,s,meas X ns,meas (3.23g) Molar flow rate of CO 2 gas at sweep outlet: hCO 2 ,s,out = XC02,s,meas X Ths,meas (3.23h) Molar flow rate of CO gas at sweep outlet: nCO,s,out = XCO,s,meas X ns,meas (3.23i) Molar flow rate of H 2 gas at sweep outlet: nH2,s,out = XH 2 ,smeas X fts,meas (3.23j) Molar flow rate of N 2 gas at sweep outlet: hN 121 2 ,s,out = XN 2 ,s,meas X fts,meas (3.23k) Constraints: 'CO 2 (3.24a) (3.24b) < hC0 2 ,s,in ,sot nCH ,soUt > hCH4 ,s,in 4 With all of the above considerations, equations are formulated for the desired variables. The oxygen permeation rate across the membrane is given by = nO2 , sperm (3.25) A J02 where A (constant at 58cm 2 ) is the membrane area exposed to gases on either side. The methane conversion (XcH4 ) is given by ?iCH4 ,s,out XCH4 X 100% (3.26) hCH4 ,s,in / The yields (Y) and selectivities (Si) of CO, C2H4, C2H and H2, can be calculated individually from Y ni x ' o x 100% (3.27) 100% (3.28) nCH4 ,s,in and Si = 2 XCH4 x where ni is the number of carbon atoms in a molecule of product i. For H2, nH 2 represents the ratio of number of hydrogen atoms in H2 to the number of hydrogen atoms in CH . 4 Therefore, nH 2 equals 1/2. 122 The yield (YC0 2 ) and selectivity (Sc0 YCO2 = nC0 2 2 ) of CO 2 are calculated from ('nco2,St n - hCO 2 ,s,in) x 100% (3.29) CH4,s,in and SCx2 3.6 YC2 X 100% (3.30) XCH4 Conclusions An ITM reactor has been presented for use in conducting oxygen permeation and methane oxymel combustion experiments. Its current purpose is to investigate the effects of operating parameters on the macroscopic thermochemical processes which occur during experiments. The design is based on a stagnation flow configuration to enable studies at the fundamental level and cross-validation with numerical work. Reactor sealing and testing mechanisms have been developed to mitigate and reduce leakage and quantify any leaks that may occur during experiments. The membrane used for the ITM reactor is Lao.9 Cao.1FeO 3-6 (LCF). It is known to provide a balance between favorable oxygen permeation rates and stability under reactive conditions [82]. Process control and instrumentation systems have been developed for the ITM reactor to enable proper control and monitoring. There are five major sections - plumbing, instrumentation, gas chromatography, heating, and safety. The plumbing network has been designed to ensure ease of flow operations, while the instruments used are intended to provide pressure, temperature and oxygen concentration readings within the reactor. A gas chromatograph 123 provides the concentrations of product gases at the reactor exhausts, while two heating systeins, enclosure and cartridge, ensure uniform and fine control of heating respectively within the reactor setup. Because of the risks involved in using methane and operating at high temperatures, a safety system has been developed for the methane flow and heating systems within the ITM reactor. During experiments, operating parameters are varied to obtain the primary variables of interest - oxygen flux and carbon dioxide yield. The experimental methodology needed to obtain these variables covers the development of reactor-centric flow procedures, heating procedures and analysis equations. The flow and heating procedures enable the variation of operating parameters while the analysis equations enable the derivation of the variables of interest. 124 Chapter 4 Results and Analysis The ITM reactor described in section 3.1 was used to carry out experiments involving variations of temperature, inlet sweep flow (mass transfer) and fuel inlet fraction (reactive analysis). This chapter covers the results and analysis of the experiments conducted. As explained in section 3.4, the performance indices are mainly oxygen permeation flux and carbon dioxide yield (oxymel combustion). In order for ITM separation methods to be competitive with current oxygen separation processes, oxygen permeation flux must be at least 7.3 tmolcm--2s1 [112,113]. Also, CO 2 capture will be a relatively easy task if only CO 2 and H2 0 are present in the product stream (since the H2 0 can be condensed). The ideal C0 2 /H 2 0 product ratio under stoichiometric conditions is 0.5. With these considerations in mind, the results presented in this chapter will provide an insight into the suitability of the ITM reactor for competitive oxygen separation and oxymel combustion. Table 4.1 shows the experimental points considered. 125 Table 4.1: Experimental points considered for analysis Experiment Controlled Parameter Permeation-only Vs,in (300sccm) T ("C) (OC)0 C 700 750 0 C 800 0 C 800 0 C 800 0 C V7 (400secmr) Qnor 500 W 40 - 800 0 C 500 500 500 500 500 500 10 15 20 T (700 0 C) T (750-C) T (800-C) Vs,in (200sccm) s,in Reactive 4.1 VCH 4 ,s,in (10sccm) VcH 4 ,s,in (15sccm) VCH 4 ,,,j (20sccm) 8000 C 8000 C Vfeed,in sweep,in (sccm) 500 500 500 500 500 (sccm) 500 500 500 200 300 VCH 4 ,s,in (sccm) Temperature 10-2 5.2 E 0 E 0 E 5.4 0 E 0 0 C 0) 5.6- 0) 0j 5.8- 0 10-3 700 750 Temperature, T [0 C] -A 800 0.95 1.05 1000/T [K- ] Figure 4-1: Dependence of oxygen permeation flux on membrane temperature (non-reactive). Feed (air) flow = 500scem, sweep (GO2 ) flow = 500scem. Figure 4-la shows the influence of temperature on oxygen permeation flux. The feed and sweep inflow rates were kept constant at 500sccm each. As expected, increase in temperature from 700 - 800'C leads to a slight increase the oxygen permeation from 4.4 - 4.9 x10--3pmolcm-2s-1. It is generally believed that an increase in temperature would lead to a higher disorder 126 in the oxygen vacancies [91, 92,109, 110]. This would in turn lead to a sharp increase in oxygen permeation. Since the oxygen permeation in figure 4-la shows only a slight overall increase, it can be assumed that the disorder-order transition of oxygen vacancies does not occur between 700 - 800*C for the LCF membrane considered in this report. The relationship between oxygen permeation flux and membrane temperature is further explored in figure 4-1b (Log J0 2 vs 1000/T). An experimental relationship between oxygen flux and temperature is obtained using regression analysis (see section 2.4.1) as: Jo2 = -4.868 exp 5.188 xX RT ) mol.cm-2 s- R2 = 0.237 (4.1) The activation energy for oxygen permeation flux (based on data from T = 700 - 800-C) is therefore 51.9kJ/mol. 850 -- Feed gas temperature -+- Sweep gas temperature CU 800 7/4 CL E a) 0) 750 C. CD a) u) 700 ~0 a) 650 650 800 750 700 Heater temperature [ C] 850 Figure 4-2: Bulk temperatures within reactor feed and sweep sides Figure 4-2 shows the bulk temperatures between the feed and sweep sides of the reactor. As the reactor temperature is increased from 700 - 800*C, the difference in bulk temperatures 127 (ATulk) reduces by half its value. The change in ATlk suggests that the heat transfer resistance of the membrane (Rmem) is reduced as the membrane temperature increases. This change in ATulk could also be due to radiation effects (as the membrane emissivity decreases with temperature) on the unshielded thermocouples. 4.2 Mass Transfer The effect of sweep flow variation on oxygen permeation flux was investigated by increasing the sweep inflow from 200 - 500sccm as shown in figure 4-3. The membrane temperature and feed (air) inflow rates were kept constant at 800'C and 500sccm respectively. The oxygen permeation flux shows a steady increase as sweep inflow is increased. The relationship in this case is fairly linear. Due to the steady increase observed, it can be assumed that the mass transfer resistance is minimum at a sweep inflow rate higher than 500sccm. C.) E 0 E p C) x 0 x c x 0 1 3 100 - - 200 - - . 300 400 Sweep inflow [sccm] . - - 500 - - 600 Figure 4-3: Dependence of oxygen permeation flux on Sweep (C0 2 ) flow (non-reactive). Membrane temperature = 800C, feed (air) flow = 500sccm. 128 C-*- 0 0 2 2 6 4- 0 0 2 0 5 10 15 20 Reynolds Number (sweep), Res [-] 25 Figure 4-4: Dependence of oxygen permeation flux and sweep 02 partial pressure on the sweep Reynolds number (non-reactive). Membrane temperature = 800C, feed (air) flow 500sccm. Re = (pY) / (pDc), De = 2.24cm. Figure 4-4) shows the influence sweep mass transfer analysis. the reactor gap height (Dc = 2.24cm) is the distance from flow straightener to the membrane, and the active volume (V = 33cm 3 ) is the volume under the membrane (i.e membrane area times gap height). As shown, for a Reynolds number increase from 6.75 - 16.89 (150% increase), the oxygen permeation flux increases by 25%. Furthermore, it is obvious that oxygen permeation increases with increasing sweep flow rate (Vin) because of the simultaneous decrease in oxygen partial pressure in the sweep side (Pf2 ). Since Jo 2 increases at a slower rate than Es,in, the bulk P 2 decreases as a result. Furthermore, the increase in flux with increasing sweep inflow suggests that the local P 2 also decreases. Since an increase in sweep mass flow rate means more oxygen is being swept away from the membrane surface, the chemical potential gradient across the membrane increases, and oxygen permeation is improved as a result [114]. 129 4.3 Reactive Analysis The results analyzed so far in this chapter are for permeation-only experiments. When fuel (CH 4 ) is added to the sweep side inflow, it reacts with permeated oxygen to form products such as C0 2 , CO, H2 0, and H2 . 4.3.1 Effect of Fuel Inlet Fraction Figure 4-5 shows the dependence of oxygen permeation flux on CH 4 inlet concentration. The membrane is kept at 800*C while the feed and sweep inflow rates are each 500sccm. As the CH 4 inlet concentration is increased from 1.96 - 3.85%, the oxygen flux increases rapidly to 22.8 x 10-3 tmolcm-2s1. This oxygen flux value is almost 400% more than that observed in the non-reactive case at the same operating conditions (although 10-20sccm CH 4 has been added to the sweep inflow in the reactive case). 10. 10 7 9 --- 2 CH4:02 100 E U E CO = 0 10 -2 8- 0 Tzzzz~ 7- C). ______________ C) 0 a.0 10-31.5 2 2.5 3 3.5 CH 4 inlet conc. [vol. %] 4 (a) 6 5 1.5 2 2.5 3 3.5 CH4 inlet conc. [vol. %] 4 (b) Figure 4-5: Dependence of (a) oxygen permeation flux; (b) sweep oxygen partial pressure, and fuel/0 2 ratio; on fuel inlet concentration. Membrane temperature = 800C, CO 2 inflow = 500sccm, feed (air) flow = 500sccm. Figure 4-5 also shows that the sweep side bulk oxygen partial pressure is quite low. This is 130 caused by reaction of oxygen with methane leading to an increased oxygen chemical potential gradient across the membrane and an increase in oxygen permeation flux. The slight decrease in oxygen partial pressure and overall increase in fuel/0 2 ratio (molar ratio of inlet CH 4 to permeated oxygen) also suggest that an increase in fuel inflow rate may be beneficial to oxygen permeation. However, since the chemical reaction rate is finite in this case, further increase in the fuel inflow rate (at CO 2 = 500sccm) would reduce the reactant residence time and further increase the fuel/0 2 ratio, thereby causing low fuel conversion [115-117]. 100 d.80 _5 G 60 0 Co 4 2 SH20 aI) Cl) L- S 40 2 202 0 o X CH -A- CSco 8 e 20- 1.5 2 2.5 3 3.5 4 4.5 5 CH4 inlet conc. [vol. %] Figure 4-6: Dependence of fuel conversion and species selectivities on fuel inlet concentration. Membrane temperature = 800'C, CO 2 inflow = 500sccm, feed (air) flow = 500sccm. Figure 4-6 shows the dependence of methane conversion and the selectivities of CO 2 , CO, H2 0, and H2 on the methane inlet concentration. The methane conversion is low and it decreases with increasing fuel inlet concentration (and fuel inflow rate). One possible reason for the low methane conversion is depicted in figure 4-7. It is possible that most of the fuel entering the reactor sweep side flows to the reactor 131 02 present in exit stream because of finite rate 02 flux IT Dc Reaction Zone Part of sweep inflow which does not see02 mixing/active region CH 4, CO2 Figure 4-7: Possible interaction of fuel with permeated oxygen in the reactor sweep side exit without reaching the reaction zone, leading to the presence of residual oxygen (about 7% of sweep outflow) in the sweep side even at high fuel/0 2 ratios where complete oxygen conversion is expected as shown in figure 4-5b. As seen in the plot, the bulk oxygen partial pressure at the reactor exit remains fairly constant despite the increase in fuel inlet flow rate beyond the stoichiometric fuel/0 2 ratio (0.5). Furthermore, an in-house 2-D numerical analysis of the reactor' (using air in both sides) shows that only about 1 - 12.5% of the total sweep inflow is advected towards the reaction zone (approximately 1 - 5mm below the membrane) as shown in table 4.2. This result shows that the primary cause of low methane conversion is the low amount of fuel which reaches 'The analysis was carried out by Prof. R. Ben-Mansour of KFUPM. 132 the reaction zone. Also, we can assume from figure 4-7 and the results in table 4.2 that a wider membrane would lead to more fuel being converted, since more fuel will reach the oxygen being permeated across the membrane. Table 4.2: Mass flow rates at different planes below the membrane (inlet sweep velocity lcm/s). Results obtained from 2-D numerical analysis of the reactor Integrated mass flow rate(kg/s) Plane location below membrane (mm) The fuel-0 2 1 5.25 3 1.90 x10- 5 Fraction of total flow (%) 0.12 x10-7 5 4.25 5.59 x10-5 12.5 10 1.82 x 10-4 40.7 15 3.10 x10-4 69.33 reaction favors H2 0 over H2 formation as shown in figure 4-6. The selec- tivity of CO steadily approaches 100% while that of CO 2 approaches 50% as the fuel inlet concentration reaches 3.85%. However this is likely due to the conversion of residual carbon (formed earlier at lower fuel inlet concentrations) in the reactor sweep side or the very fuel-rich nature of the reactor at that point. Furthermore, increased CO formation could be due to the nature of the methane reaction kinetics. As discussed earlier, the fuel/0 2 ratio in this case is already high (fig 4-5) meaning there is more fuel than required for complete oxymel combustion. Since the kinetics of methane oxycombustion (equation 2.10) is extremely fast, it is possible that the remaining CH 4 not consumed by the permeated oxygen may react with CO 2 producing H2 and CO via equation 4.2 below [93]: CH 4 + CO 2 a 2CO + 2H2 133 AH29 8 247 kJ/mol (4.2) 4.3.2 Oxymel Analysis Figure 4-8 shows the dependence of CO 2 yield and C0 2 :H2 0 product ratio on the reactor fuel input. From the plot, it can be seen that the ITM reactor is not producing a high amount of oxymel products at the operating reactor conditions. The overall CO 2 yield is much lower than the desired value of 100%. The yield increases with fuel inflow, but the highest value obtained during reactive experiments at 800'C is 2.6%. The low yield can however be attributed to the fact that not much of the fuel reaches the membrane/reaction zone. 3 . ., E3YCo 2 - 2.5 - CO2:H20 2 2 0 2-r (N 0 0 1.5 0 (N0.5 0 5 _ . . . . . ._ 10 _. _ 15 20 CH inflow rate [sccm] . . . . .. 25 Figure 4-8: Analysis of oxymel products obtained from reaction between fuel and 02 in the ITM reactor. Membrane temperature = 800*C, CO 2 inflow = 500sccm, feed (air) flow = 500sccm. However, the C0 2 /H 2 0 ratio is not far from ideal oxymel reaction regimes (C0 /H 0 2 2 = 0.5). However, this ratio is more likely to be achieved with reduction in the fuel inflow. The obvious disadvantage of the fuel increase is a simultaneous reduction in fuel conversion. 134 Also, the fuel/0 ratio would increase further leading to the potential formation of more CO 2 and H2 via equation 4.2. 4.3.3 Influence of Reactive (reducing) Sweep Gas on Oxygen Flux 10 Cn E 0 E 0 10 -2 0) X 0 10 0 5 15 10 CH4 inflow rate [sccm] 20 Figure 4-9: Influence of fuel addition on oxygen permeation flux in the ITM reactor. Membrane temperature 800 0C, feed (air) flow = 500sccm, sweep flow = 500sccm CO 2 + variable CH 4. As shown in figure 4-9, the flux observed in the reactive experiments can be an order of magnitude higher than that observed in the non-reactive case. As already discussed, this is mainly due to the further decrease in the sweep side partial pressure of oxygen due to fuel consumption in the reactive case. It is fair to assume that much higher fluxes will be observed when sweep inflow rates are increased at higher temperatures (> 800*C) as widely reported in literature [15,50,77,101, 118-120]. As seen from the oxygen flux trends in figure 4-9, the reactive flux is likely to 135 show steady increase with further increase in temperature as reported in literature. 4.4 Comparison with other ITM Experimental Investigations The experimental results analyzed so far in this chapter are reflective of the initial operations of the ITM reactor; the reactor is yet to be operated beyond a membrane temperature of 800*C and inflow rates beyond 500sccm. However, a comparison with results from state of the art ITM reactors can provide an insight into the regimes within which the ITM reactor would operate best thereby aiding the reactor optimization process. In this section, comparisons are made with the state of the art ITM results discussed in section 2.4. Oxygen permeation fluxes have been normalized for comparison as explained in section 2.4.1. 4.4.1 Temperature Comparison In figure 4-10, a comparison is made with oxygen permeation fluxes reported in literature. The ITM reactor currently produces lower flux than most investigations in literature. However, the membrane used (LCF) is more suitable for oxygen permeation under reactive conditions. Also, effects of temperature beyond 800'C are yet to be explored. Most of the membranes in figure 4-10 are only stable for oxygen permeation and other non-reactive processes. The reports which utilize reactive-capable membranes (e.g Lu 2000) show similar or slightly higher fluxes compared to those reported in this study. Figure 4-11 shows normalized flux as a function of temperature for non-reactive cases. The plot shows 136 1 ----Zeng 1998 1 )1(-- Lu 2000 Zeng 2000 Zeng 2001 A -X- 0-- -\7-Tong 2002 (a) Tong 2002 (b) -1 -_-<}- Diethelm 2004 (a) -*-Wang 2004 (a) oE *-- -2 Ikeguchi 2005 (a) Liu 2005 (a) *---- E - 2 o 0) Liu 2006 (a) Liu 2006 (e) Liu 2006 (i) ---- 3 Wang 2005 (a) 4Zhang 0 2007 (a) 2007 (c) 4--Zhang Buysse 2009 (a) 5~ . ' -6 0.7 Diethelm 2009 (a) - ' ' ' 0.8 . 0.9 . 1 1000/T [K l] 1.1 ' 1.3 -- 1.2 Wei 20 (a) *Zhu 2010 Shen 2011 (b) This Study Figure 4-10: Oxygen flux dependence on temperature (non-reactive) for the ITM reactor compared with reported investigations. -+- Zeng 1998 -*- Lu 2000 - -X--Zeng 2000 Zeng 2001 --V-Tong 2002 (a) -V-Tong 2002 (b) 1.4 1.31.2- -<- Diethelm 2004 (a) Wang 2004 (a) Ikeguchi 2005 (a) Liu 2005 (a) -*-Wang 2005 (a) 7P 1.1 ---- ~-Y1 - .- Liu 2006 (a) Liu 2006 (e) --- * 0-y-- Liu 2006 (i) -4- 0.9 -- -*- 0.8-* 0.7 550 Zhang 2007 (a) 4- Zhang 2007 (c) ' 600 650 700 750 800 850 Temperature, T rc] ' 900 ' ' 950 ' ' 1000 Buysse 2009 (a) Diethelm 2009 (a) Wei 2010 (a) 2010 --*Zhu - . 105 -- Shen 2011 (b) This Study Figure 4-11: Normalized oxygen flux dependence on temperature (non-reactive) for the ITM reactor compared with reported investigations. See section 2.4.1 for normalization methodology. the normalized flux as a function of temperature to be close to 1. Therefore, the fluxes obtained from the ITM reactor in this study can be normalized using the methods described earlier in section 2.4.1. 137 4.4.2 Mass Transfer Comparison In terms of mass transfer, the ITM reactor operates similar to the reported investigations for non-reactive cases as seen in figure 4-12. The flow Reynolds number is within similar range and the oxygen flux trend indicates that the mass transfer resistance is close to its minimum value at the operating membrane temperature (800C). An analysis at lower inflow rates is more likely to reveal more details at this temperature. Wang 2004 (e) Wang 2004 (f) 5-.A5- .....2-- -2---*- .. -- 7 -- - --- 1.5 -'1 . .. . - 0.5*- ''-Wei 0 2 4 6 8 10 ReRsweep [-- 12 14 Liu 2006 (c) -Liu 2006 (g) Liu 2006 (h) Liu 2006 (j) -v- Liu 2006 (k) --- 0 Wang 2004 (g) Liu 2005 (b) Liu 2005 (c) Liu 2005 (d) Liu 2006 (b) Liu 2006 (d) Liu 2006 (f) .--- 1.5 - - 16 18 --- Liu 2006 (1) Buysse 2009 (b) Wei 2010 (g) Wei 2010 (h) Wei 2010 (i) Wei 2010 (j) 2010 (k) This Study Figure 4-12: Normalized flux dependence on sweep mass flow (non-reactive) for the ITM reactor compared with reported investigations. See section 2.4.1 for normalization methodology. A comparison of the sweep flow timescale as shown in figure 4-13 reflects the difference in the sweep residence time compared with reported investigations. Since the ITM reactor in this study has an active volume (V = 33cm 3 ) which is about 1-2 orders of magnitude above the reported investigations, this result is expected. Also, all the experiments in figure 413 involve tubular or hollow fiber membrane geometries. A direct comparison with similar configurations (disk ITM reactors) to the ITM reactor in this study would be ideal. However, 138 # -4-- Wang 2004 (e) 2.5 . . . ', . . . Liu 2005 (c) 2@- --7 S0" 1.5 >-y- LL-y- V 1 S -..- Liu 2005 (d) Liu 2006 (b) Liu 2006 (c) - Liu Liu Liu Liu 2006 (d) 2006 (f) 2006 (g) 2006 (h) Liu 2006 (j) -v-- Liu 2006 (k) Z - 0.5 0 10 (f) Wang 2004 (g) -Liu 2005 (b) - 5 Wang 2004 10 10 Residence Time (sweep) [s] 10 101 Liu 2006 (1) Buysse 2009 (b) Wei 2010 (g) Wei2010(h) Wei 2010 (i) Wei 2010 (j) Wei2010(k) This Study Figure 4-13: Normalized flux dependence on Residence time (non-reactive) for the ITM reactor compared with reported investigations. 7-res = V/V. See section 2.4.1 for normalization methodology. the active volume of these setups are difficult to estimate as the geometries are generally not listed in literature. 4.4.3 Reactive Comparison From figure 4-14, it can be seen that the fuel conversion in the ITM reactor is below equilibrium. However, the trend suggests that a reduction in fuel/0 2 ratio may favor increased fuel conversion and quasi-equilibrium operation of the ITM reactor. As discussed earlier, the low fuel conversion could also be due to most of the inlet methane leaving the reactor sweep side without reaching the reaction zone (figure 4-7). However, while high fuel conversion is important, it should be noted that the ITM reactor in this study was not designed specifically for high reactive performance, but for spatially resolved measurements. 139 100 Dong 2001 (c) Dietheg 200() 80 09200 () Wag2005a) 70Tan 200() 60 - (1) a 2 T10620086(b) 20 C Tano 2006 (f) -20064) L) .0 40 Fu > 41 A) n Zh22000(b) 2008 008 , Kozh-o'20009( ) L.02010)0) 10T1n TZ L4o2010b2 30 on) o200e T4020,0(c) C.Gon 4.5 1420116() 20 Gong 201() KnLe0 201 () L00a2011(b) Googp 2011 (c) Concl20us(i) CH 20 : 02 ratio[-] Knoeg 2011 (0) * 0 510 *h 15 Thu-tod CH 4 :02 ratio[- Figure 4-14: Dependence of CH4 conversion on fuel/0 with reported investigations. 4.5 2 ratio for the JTM reactor compared Conclusions Preliminary experiments have been conducted using the ITM reactor described in section 3.1. Non-reactive experiments were conducted by varying the membrane temperature and the sweep inflow rate of CO2. Reactive analyses of the effects of fuel inlet fraction were also carried out. Furthermore, comparisons were made with reported experimental investigations in literature. When the membrane temperature was varied under non-reactive conditions, an overall increase in oxygen flux was observed up to 4.9 x 10- 3 pmolcm- 2s-1 at 800*C. The activation energy required for oxygen permeation across the membrane was obtained as 51.9kJ/mol. Further analysis shows that the difference in bulk temperatures between the feed and sweep gases decreases with increasing temperature. This suggests that the heat transfer resistance across the LCF membrane decreases with temperature or the decreasing membrane emissivity 140 (with increasing temperature) affects the unshielded thermocouple readings. The effect of increasing flow rate in the reactor sweep side was studied to provide insight into its impact on oxygen permeation flux. Oxygen flux shows an increase with sweep mass flow mainly because of a simultaneous decrease in sweep side oxygen partial pressure. Even though the increase is not rapid, it can be assumed that the mass transfer resistance is yet to reach its minimum value. An analysis of the sweep gas timescale reveals that a reduction in residence time favors oxygen permeation. A study of the effects of fuel addition to sweep inflow reveals that oxygen permeation is enhanced beyond that observed in the non-reactive case. However, the sweep side partial pressure of oxygen in this case decreases gradually suggesting that most of the fuel entering the reactor sweep side does not reach the reaction zone. In addition, the increase in fuel flow rate increases the fuel/0 2 ratio. These factors are likely causes for the low fuel conversion observed. From the species formation values, it is observed that the reactions are favorable toward CO and H2 0 rather than CO 2 and H2 respectively. Although, in the case of CO formation, residual CH 4 may interact with CO 2 to form more CO as the fuel/0 2 ratio increases in the reactor sweep side. The mass transfer results observed in this study are in agreement with reported data; the flux increases gradually with mass inflow. However, the residence time within the reactor was about 2 orders of magnitude higher than reported in literature, due to the correspondingly large active volume of the ITM reactor. The fuel conversion is low compared to reported data. However, there is an indication that lower fuel/0 2 ratios will favor fuel conversion. Overall, the maximum flux observed in this study (22.8 x10-3pmolcm-2s-') is about 2 orders of magnitude less than required to compete with current oxygen separation technolo141 gies. However, this flux is observed at a relatively low temperature (800C) compared to reported investigations. According to the trends observed in literature, the flux is likely to increase appreciably if temperature is increased up to 950C. In terms of the oxymel operating capability of the ITM reactor, it is fair to assume that the desired C0 2 :H2 0 can be achieved at the current operating temperature of 800*C, but high CO 2 selectivity is more likely to occur with a lower fuel/0 2 ratio (= 0.5) at higher temperatures and with a more appropriate flow configuration. 142 Chapter 5 Conclusions This work provides a report on the development of an ion transport membrane reactor for use in oxygen separation from air and combustion with methane. The reactor was used for experiments involving analysis of the effects of temperature, mass flow, and fuel on oxygen permeation and other thermochemical processes. An analysis of reported experimental investigations in literature was also presented. 5.1 Summary The design details and analysis of initial operation of a novel ITM reactor have been presented. The reactor operates with bi-directional stagnation flow and can support optical analysis of the reaction zone and spatial analysis along the reactor axes. The application of the reactor is to investigate oxygen permeation across a planar membrane and facilitate combustion of the permeated oxygen with fuel. Sealing methodologies have been developed and an effective strategy has been presented. 143 The membrane used for the ITM reactor experiments was La 0 .9 Cao. 1 FeO 33 (LCF). Preliminary experiments conducted show that the maximum oxygen flux obtained at a membrane temperature of 800'C is 2 orders of magnitude less that is required to compete with current oxygen separation technologies. However, the trends reported in literature suggest that much higher fluxes can be obtained at higher temperatures. The conversion of methane and overall CO 2 yield are also observed to be below low ( < 10%). However, improvements may be observed if the fuel inlet flow rate is reduced at higher temperatures, and a more appropriate flow configuration is used. 5.2 Outlook and Future Work The work presented in this reported covers the early operation of the ITM reactor. Further analyses, improvements, and modifications can still be obtained with the ITM reactor. The next steps for the ITM reactor include further global measurements and spatial measurements. 5.2.1 Global Measurements In order to further understand the capabilities of the ITM reactor, it is important to carry out further global experiments. From the discussion in chapter 4, it is evident that certain operating regimes are yet to be explored. The reactor is yet to be operated beyond 800'C. It is safe to assume based on reports in literature, that more improved results for oxygen flux, fuel conversion and CO 2 yield can be obtained at higher temperatures. Also, it may be possible to adjust other conditions such as the reactor gap height and flow rates to 144 Table 5.1: Experimental points for future consideration Experiment T (OC) (sccm) VCo 2,in (sccm) (700-C) (750-C) (800-C) (850-C) (900-C) T (950-C) V,in (100sccm) Vs8in (200sccm) V8,in (300sccm) VSin (400sccm) 700 750 800 850 900 950 950 950 950 950 500 500 500 500 500 500 500 500 500 500 500 500 500 500 500 500 100 200 300 400 V 8 ,in (500sccm) 950 500 500 700 750 800 850 900 950 950 950 950 950 950 950 950 950 950 950 950 500 500 500 500 500 500 500 500 500 500 500 500 500 500 500 500 500 450 450 450 450 450 450 90 180 270 360 450 500 490 480 470 460 450 Controlled Parameter T T T T T Permeation-only (700-C) (750-C) (800-C) (850-C) (900-C) (950-C) V,in (1OOscem) (200sccm) QV'in (300sccm) V,in (400sccm) V,in (500sccm) XCH 4 ,s,in (0%) XCH 4 ,s,in (2%) XCH 4 ,s,in (4%) XCH 4 ,s,in (6%) XCH 4 ,s,in (8%) XCH 4 ,s,in (10%) V,in Reactive Vairin VCH 4 ,s,in (sccm) 10 20 30 40 50 obtained improved results. If a wide range of data is obtained, direct comparisons can be made between reactive and non-reactive conditions based on variations in temperature and flow rate. Furthermore, the effects of the reactive sweep gas can be better understood if the total sweep gas flow and fuel dilution ratio are varied separately (keeping every other condition constant each time). 145 Table 5.1 can be used to carry out a wide range of measurements which would produce enough data for a more complete analysis of the ITM reactor. 5.2.2 Spatial analysis Based on the importance of localized effects within the ITM reactor (such as local P0 2 ), the use of optical measurement techniques will aid the study of species formation in the reaction zone and its controlling parameters (e.g. flow and temperature). Optical equipment such as a Scanning Infrared Gas Imaging System could be placed outside of the reactor and aimed through the optical access to detect gas species within the reaction zone. Importantly, the reactive oxygen species within the reaction zone can be monitored using the hydroxylradical detection method. As a result, oxygen permeation and flow gradient (within the sweep side can be analyzed). Also, the gas sampling points along the reactor length can be used to study the effects of species, temperature and pressure gradients within the reactor. As discussed in section 3.1, GC measurement and other sensing probes can be placed across the length of the reactor. The measurements obtained can be compared with the values at the reactor exit. Importantly, the local oxygen partial pressures on both sides of the reactor can be obtained leading to a fundamental understanding of the oxygen flux governing processes. This would also aid the understanding of the relative impacts of surface exchange and bulk processes on oxygen permeation and facilitate cross-validation with in-house numerical work carried out on the ITM reactor. 146 Bibliography [1] Hanne M. Kvamsdal, Kristin Jordal, and Olav Bolland. A quantitative comparison of gas turbine cycles with CO 2 capture. Energy, 32(1):10-24, 2007. [2] Burggraaf A.J. Bouwmeester H.J.M. Dense Ceramic membranes for Oxygen Separation in: CRC Handbook of Solid State Electrochemistry, chapter 14. CRC Press, Inc., 1997. [3] Crystal Lattice Structures. Center for Computational Materials Science of the United States Naval Research Laboratory (17 Feb 2007). Retrieved July 21, 2011 from http://cst-www.nrl.navy.mil/lattice/. [4] Timothy Griffin, Sven Gunnar Sundkvist, Knut Asen, and Tor Bruun. Advanced zero emissions gas turbine power plant. Journal of Engineeringfor Gas Turbines and Power, 127(1), 2005. [5] Yutie Liu, Xiaoyao Tan, and K. Li. Mixed conducting ceramics for catalytic membrane processing. Catalysis Reviews: Science and Engineering, 48(2):145 - 198, 2006. [6] J. Sunarso, S. Baumann, J. M. Serra, W. A. Meulenberg, S. Liu, Y. S. Lin, and J. C. Diniz da Costa. Mixed ionic-electronic conducting (MIEC) ceramic-based membranes for oxygen separation. Journal of Membrane Science, 320(1-2):13-41, 2008. [7] W. Wang and Y. S. Lin. Analysis of oxidative coupling of methane in dense oxide membrane reactors. Journal of Membrane Science, 103(3):219-233, 1995. [8] Weishen Yang and Rui Cai. Oxygen-Ion Transport Membrane and Its Applications in Selective Oxidation of Light Alkanes in: Inorganic Membranes for Energy and Environmental Applications, chapter 14. Springer, Pittsburgh, PA, USA, 2009. [9] Henny J. M. Bouwmeester. Dense ceramic membranes for methane conversion. Catalysis Today, 82(1-4), 2003. [10] Robert G. Watts. Global warming: Is it real? Journal of Solar Energy Engineering, 122(3):158-160, 2000. [11] Projections: EIA, World Energy Projection System Plus (2010). EIA, International Energy Statistics database (as of November 2009). Retrieved July 6, 2011 from http://www.eia.gov/oiaf/ieo/emissions.html. 147 [12] Massachusetts Institute of Technology. MIT, Cambridge MA, 2010. The Future of Natural Gas [Interim report]. [13] Yosuke Takahashi, Akihiro Kawahara, Takehiro Suzuki, Masayoshi Hirano, and Woosuck Shin. Perovskite membrane of LajxSrxTijyFeyO3-deita for partial oxidation of methane to syngas. Solid State Ionics, 181(5-7):300-305, 2010. [14] Chun Zhang, Xianfeng Chang, Xueliang Dong, Wanqin Jin, and Nanping Xu. The oxidative stream reforming of methane to syngas in a thin tubular mixed-conducting membrane reactor. Journal of Membrane Science, 320(1-2), 2008. [15] Xiaoyao Tan, K. Li, A. Thursfield, and I. S. Metcalfe. Oxyfuel combustion using a catalytic ceramic membrane reactor. Catalysis Today, 131(1-4):292-304, 2008. [16] Clark D.J. Suitor J.W. and Losey R.W. Development of alternative oxygen production source using a zirconia solid electrolyte membrane, in: Technical progress report for fiscal years 1987, 1988 and 1990, Jet Propulsion Laboratory Internal Document D7790, 1990. [17] S. S. Liou and W. L. Worrell. Electrical properties of novel mixed-conducting oxides. Applied Physics A: Materials Science &amp; Processing,49(1):25-31, 1989. [18] Worrell W.L. Liou, S.S. Mixed conducting electrodes for solid oxide fuel cells. Proc. 1st Int. Symp. on Solid Oxide Fuel Cells, Singhal, S.C., Ed., The Electrochemical Society, Pennington, NJ, pages 81-89, 1989. [19] B. Cals and J.F. Baumard. Electrical properties of the ternary solid solutions. 17:137147, 1980. [20] B. Cales and J. F. Baumard. Mixed conduction and defect structure of ZrO CeO 2 2 Y2 0 3 solid solutions. Journal of The Electrochemical Society, 131(10):2407-2413, 1984. [21] T. J. Mazanec, T. L. Cable, and J. G. Frye. Electrocatalytic cells for chemical reaction. Solid State Ionics, 53-56(Part 1):111 - 118, 1992. [22] W. C. Maskell and B. C. H. Steele. Solid state potentiometric oxygen gas sensors. Journal of Applied Electrochemistry, 16:475-489, 1986. 10.1007/BF01006843. [23] Zhentao Wu, Bo Wang, and K. Li. A novel dual-layer ceramic hollow fibre membrane reactor for methane conversion. Journal of Membrane Science, 352(1-2):63-70, 2010. [24] Alexander S. Mukasyan, Colleen Costello, Katherine P. Sherlock, David Lafarga, and Arvind Varma. Perovskite membranes by aqueous combustion synthesis: synthesis and properties. Separation and Purification Technology, 25(1-3):117 - 126, 2001. [25] Paul W. 0. Hoskin. Minerals: Their constitution and origin: by Hans-Rudolf Wenk and Andrei G. Bulakh. Cambridge University Press, Cambridge, U.K., 2004. 646 pp., 526 ill + 16 plates. American Mineralogist, 91(7):1210, 2006. 148 [26] Kenneth W. Bladh John W. Anthony, Richard A. Bideaux and Eds. Monte C. Nichols. Handbook of Mineralogy, Mineralogical Society of America. Chantilly, VA 20151-1110, USA. [27] L. L. Hench and L. K. West. Principles of Electronic Ceramics. John Wiley & Sons, Inc., 1990. [28] I. D. Brown. What factors determine cation coordination numbers? Acta Crystallographica Section B, 44(6):545-553, Dec 1988. [29] Zhang H.M. Furukawa S. Teraoka, Y. and N. Yamazoe. Oxygen permation through perovskite-type oxides. Chem. Lett., pages 1743-46, 1985. [30] Nobunaga T. Teraoka, Y. and N. Yamazoe. Effect of cation substitution on the oxygen semipermeability of perovskite oxides. Chem. Lett., pages 503-06, 1988. [31] Y. Teraoka, T. Nobunaga, K. Okamoto, N. Miura, and N. Yamazoe. Influence of constituent metal cations in substituted LaCoO 3 on mixed conductivity and oxygen permeability. Solid State Ionics, 48(3-4):207-212, 1991. [32] J. G. McCarty and H. Wise. Perovskite catalysts for methane combustion. Catalysis Today, 8(2):231-248, 1990. [33] Y. S. Lin and Y. Zeng. Catalytic properties of oxygen semipermeable perovskite-type ceramic membrane materials for oxidative coupling of methane. Journal of Catalysis, 164(1):220-231, 1996. [34] Junichiro Mizusaki, Yasuo Mima, Shigeru Yamauchi, Kazuo Fueki, and Hiroaki Tagawa. Nonstoichiometry of the perovskite-type oxides LapxSrxCoO 3-deita. Journal of Solid State Chemistry, 80(1):102-111, 1989. [35] Xiaoyao Tan and K. Li. Design of mixed conducting ceramic membranes/reactors for the partial oxidation of methane to syngas. AIChE Journal, 55(10):2675-2685, 2009. [36] Dixon AG in:. Spivey JJ, editor. Catalysis. Vol. 14. London: Royal Society of Chemistry, 1999. [37] Yaping Lu, Anthony G. Dixon, William R. Moser, Yi Hua Ma, and Uthamalingam Balachandran. Oxidative coupling of methane using oxygen-permeable dense membrane reactors. Catalysis Today, 56(1-3):297-305, 2000. [38] Haihui Wang, You Cong, and Weishen Yang. Oxidative coupling of methane in Bao.5 Sro.5 Co0 .Fe 2 O3_ tubular membrane reactors. Catalysis Today, 104(2-4):160 167, 2005. Catalysis in Membrane Reactors. [39] McGovern J. Foy, K. Comparison of ion transport membranes. Fourth Annual Conference on carbon capture and sequestration DOE/NETL, 2005. 149 [40] Y. Zeng, Y. S. Lin, and S. L. Swartz. Perovskite-type ceramic membrane: Synthesis, oxygen permeation and membrane reactor performance for oxidative coupling of methane. Journal of Membrane Science, 150(1):87-98, 1998. [41] T. Takahashi and Denki Kagaku Iwahara, H., 1967. [42] T. Takahashi and H. Iwahara. Ionic conduction in perovskite-type oxide solid solution and its application to the solid electrolyte fuel cell. Energy Conversion, 11(3):105 111, 1971. [43] B. C. H. Steele. Oxygen ion conductors and their technological applications. Materials Science and Engineering: B, 13(2):79-87, 1992. [44] Elshof J. E. Ten, H. J. M. Bouwmeester, and H. Verweij. Oxidative coupling of methane in a mixed-conducting perovskite membrane reactor. Applied Catalysis A:General, 130(2):195-195, 1995. [45] Y. Zeng and Y. S. Lin. Oxygen permeation and oxidative coupling of methane in yttria doped bismuth oxide membrane reactor. Journal of Catalysis, 193(1):58-64, 2000. [46] Y. Zeng and Y. S. Lin. Oxidative coupling of methane on improved bismuth oxide membrane reactors. AIChE Journal,47(2):436-444, 2001. [47] U. Balachandran, J. T. Dusek, R. L. Mieville, R. B. Poeppel, M. S. Kleefisch, S. Pei, T. P. Kobylinski, C. A. Udovich, and A. C. Bose. Dense ceramic membranes for partial oxidation of methane to syngas. Applied Catalysis A: General, 133(1):19-29, 1995. [48] Masayuki Ikeguchi, Tomohiro Mimura, Yasushi Sekine, Eiichi Kikuchi, and Masahiko Matsukata. Reaction and oxygen permeation studies in Sm 0 .4 Bao.6 Fe0 .sCo0 .2 0 3 - membrane reactor for partial oxidation of methane to syngas. Applied Catalysis A: General, 290(1-2):212-220, 2005. [49] Haihui Wang, Armin Feldhoff, Jurgen Caro, Thomas Schiestel, and Steffen Werth. Oxygen selective ceramic hollow fiber membranes for partial oxidation of methane. AIChE Journal,55(10):2657-2664, 2009. [50] Xiaoyao Tan and K. Li. Oxidative coupling of methane in a perovskite hollow-fiber membrane reactor. Industrial and Engineering Chemistry Research, 45(1):142-149, 2006. [51] Shokotov M. Yantovski, E. Zero emissions membrane piston engine system (ZEMPES). Proc. 2nd Annual Conference on Carbon Sequestration, 2003. [52] F. Iguchi, N. Sata, H. Yugami, and H. Takamura. Oxygen permeation properties and the stability of Lao.6 Sro.4 Feo.8 Coo.20 3 studied by raman spectroscopy. Solid State Ionics, Diffusion &amp; Reactions, 177(26-32):2281-4, 2006. [53] Nanping Xu, Shiguang Li, Wanqin Jin, Jun Shi, and Y. S. Lin. Experimental and modeling study on tubular dense membranes for oxygen permeation. AIChE Journal, 45(12):2519-2526, 1999. 150 [54] Shiguang Li, Wanqin Jin, Pei Huang, Nanping Xu, Jun Shi, and Y. S. Lin. Tubular lanthanum cobaltite perovskite type membrane for oxygen permeation. Journal of Membrane Science, 166(1):51-61, 2000. [55] Kirsten Foy and Jim McGovern. Analysis of the effects of combining air separation with combustion in a zero emissions (ZEITMOP) cycle. Energy Conversion and Management, 48(11):3046-3052, 2007. [56] C. Buysse, A. Kovalevsky, F. Snijkers, A. Buekenhoudt, S. Mullens, J. Luyten, J. Kretzschmar, and S. Lenaerts. Fabrication and oxygen permeability of gastight, macrovoidfree Bao.5 Sro.5Coo.sFeo.203-delta capillaries for high temperature gas separation. Journal of Membrane Science, 359(1-2):86-92, 2009. [57] Hui Lu, Jianhua Tong, You Cong, and Weishen Yang. Partial oxidation of methane in Bao. 5 Sro.5CoosFeo.203-delta membrane reactor at high pressures. Catalysis Today, 104(2-4):154 - 159, 2005. Catalysis in Membrane Reactors. [58] Qing Zeng and Yan-bo Zuo and Chuan-gang Fan and Chu-sheng Chen. C0 2 -tolerant oxygen separation membranes targeting CO 2 capture application. Journal of Membrane Science, 335(1-2):140 - 144, 2009. [59] Stefan Engels, Torsten Markus, Michael Modigell, and Lorenz Singheiser. Oxygen permeation and stability investigations on MIEC membrane materials under operating conditions for power plant processes. Journal of Membrane Science, 370(1-2):58 - 69, 2011. [60] F. T. Akin and Y. S. Lin. Selective oxidation of ethane to ethylene in a dense tubular membrane reactor. Journal of Membrane Science, 209(2):457-467, 2002. [61] Adrian Leo, Shaomin Liu, and Joao C. Diniz da Costa. Development of mixed conducting membranes for clean coal energy delivery. InternationalJournalof Greenhouse Gas Control, 3(4):357-367, 2009. [62] Igor Kul'bakin, Valery Belousov, Sergey Fedorov, and Anatoly Vorobiev. Solid/melt ZnO-Bi 2O 3 composites as ion transport membranes for oxygen separation from air. Materials Letters, 67(1):139 - 141, 2012. [63] E. Yantovski, J. Gorski, B. Smyth, and J. ten Elshof. Zero-emission fuel-fired power plants with ion transport membrane. Energy, 29(12-15):2077-2088, 2004. [64] Shokotov M. McGovern J. Vaddella V. Yantovski, E. Zero emissions membrane piston engine system for a bus. Proc. VAFSEP 2004, 2004. [65] Shokotov M. Shokotov V. McGovern J. Foy K. Yantovski, E. Elaboration of zero emissions membrane piston engine system (ZEMPES) for propane fuelling. Proc. 4th Annual Conference on Carbon Sequestration, 2005. 151 [66] Sven Gunnar Sundkvist, Stein Julsrud, Bent Vigeland, Tyke Naas, Michael Budd, Hans Leistner, and Dieter Winkler. Development and testing of AZEP reactor components. InternationalJournal of Greenhouse Gas Control, 1(2):180-187, 2007. [67] Sherman J. Xu and William J. Thomson. Oxygen permeation rates through ionconducting perovskite membranes. Chemical Engineering Science, 54(17), 1999. [68] C.S. Chen, B.A. Boukamp, H.J.M. Bouwmeester, G.Z. Cao, H. Kruidhof, A.J.A. Winnubst, and A.J. Burggraaf. Microstructural development, electrical properties and oxygen permeation of zirconia-palladium composites. Solid State Ionics, 76(1-2):2328, 1995. [69] G. Z. Cao. Electrical-conductivity and oxygen semipermeability of terbia and yttriastabilized zirconia. Journal of Applied Electrochemistry, 24(12):1222-1227, 1994. [70] A.J. Burggraaf P.J. Gellings H.J.M. Bouwmeester, H. Kruidhof. Oxygen semipermeability of terbia-stabilized bismuth oxide. Solid State Ionics, 460:53-56, 1992. [71] S. Dou, C. R. Masson, and P. D. Pacey. Mechanism of oxygen permeation through lime-stabilized zirconia. Journal of The Electrochemical Society, 132(8), 1985. [72] Yee Terrence F. Drake Miles P. Thorogood Robert M., Srinivasan Rajagopalan. Composite mixed conductor membranes for producing oxygen, August 1993. [73] E. Capoen, G. Nowogrocki, R. J. Chater, S. J. Skinner, J. A. Kilner, M. Malys, J. C. Boivin, G. Mairesse, and R. N. Vannier. Oxygen permeation in bismuth-based materials. Part II: Characterisation of oxygen transfer in bismuth erbium oxide and bismuth calcium oxide ceramic. Solid State Ionics, 177(5-6):489-492, 2006. [74] S. Kim, Y.L. Yang, A.J. Jacobson, and B. Abeles. Oxygen surface exchange in mixed ionic electronic conductor membranes. Solid State Ionics, 121(1-4):31 - 36, 1999. [75] M.J. Verkerk and A.J. Burggraaf. Oxygen transfer on substituted ZrO 2 , Bi 2 0 3 , and CeO 2 electrolytes with platinum electrodes I. electrode resistance by D-C polarization. Journal of the Electrochemical Society, 130:70-78, 1983. [76] M.J. Verkerk and A.J. Burggraaf. Oxygen transfer on substituted ZrO 2 , Bi 2 03, and CeO 2 electrolytes with platinum electrodes II. A-C impedance study. Journal of the Electrochemical Society, 130(1):78-84, 1983. [77] S. Liu, X. Tan, Z. Shao, and J. C. Diniz da Costa. Bao.5Sro-5 Coo.sFeo.2O3-deIta ceramic hollow-fiber membranes for oxygen permeation. AIChE Journal, 52(10):3452-3461, 2006. [78] Xiaoyao Tan and K. Li. Modeling of air separation in a LSCF hollow-fiber membrane module. AIChE Journal,48(7):1469-1477, 2002. [79] Jonas Gurauskis, rjan Fossmark Lohne, Hilde Lea Lein, and Kjell Wiik. Processing of thin film ceramic membranes for oxygen separation. Journal of the European Ceramic Society, 32(3):649 - 655, 2012. 152 [80] U. Balachandran, J. T. Dusek, P. S. Maiya, B. Ma, R. L. Mieville, M. S. Kleefisch, and C. A. Udovich. Ceramic membrane reactor for converting methane to syngas. Catalysis Today, 36(3):265-272, 1997. [81] Carl A. Udovich, D. Sanfilippo F. Frusteri A. Vaccari A. Parmaliana, and Arena F. Ceramic membrane reactors for the conversion of natural gas to syngas. In Studies in Surface Science and Catalysis, volume Volume 119, pages 417-422. Elsevier, 1998. [82] Paul N. Dyer, Robin E. Richards, Steven L. Russek, and Dale M. Taylor. Ion transport membrane technology for oxygen separation and syngas production. Solid State Ionics, 134(1-2):21-33, 2000. [83] S. Diethelm, J. Van herle, P. H. Middleton, and D. Favrat. Oxygen permeation and stability of Lao.4Cao.6FexCoxO3-de1ta (x = 0, 0.25, 0.5) membranes. volume 118 of Journal of Power Sources, pages 270-275, Amsterdam, Netherlands, 2003. Elsevier. [84] Xuefeng Zhu, Haihui Wang, and Weishen Yang. Relationship between homogeneity and oxygen permeability of composite membranes. Journal of Membrane Science, 309(1-2):120 - 127, 2008. [85] Xuefeng Zhu, Qiming Li, Yufeng He, You Cong, and Weishen Yang. Oxygen permeation and partial oxidation of methane in dual-phase membrane reactors. Journal of Membrane Science, 360(1-2):454 - 460, 2010. [86] Xuefeng Zhu, HaiHui Wang, You Cong, and Weishen Yang. Partial oxidation of methane to syngas in BaCeo.15 Feo. 5 0 3 membrane reactors. Catalysis Letters, 111(34):179-185, 2006. [87] Wenliang Zhu, Guoxing Xiong, Wei Han, and Wenshen Yang. Catalytic partial oxidation of gasoline to syngas in a dense membrane reactor. Catalysis Today, 93-95:257-261, 2004. [88] D. Schlehuber, E. Wessel, L. Singheiser, and T. Markus. Long-term operation of a Lao. 5sSro.4Coo.2Feo.sO3-delta membrane for oxygen separation. Journal of Membrane Science, 351(1-2):16-20, 2010. [89] Wenxing Zhang, J. Smit, M. van Sint Annaland, and J. A. M. Kuipers. Feasibility study of a novel membrane reactor for syngas production: Part 1: Experimental study of 02 permeation through perovskite membranes under reducing and non-reducing atmospheres. Journal of Membrane Science, 291(1-2), 2007. [90] X. Qi, F. T. Akin, and Y. S. Lin. Ceramic-glass composite high temperature seals for dense ionic-conducting ceramic membranes. Journal of Membrane Science, 193(2):185 - 193, 2001. [91] L. Qiu, T. H. Lee, L. M. Liu, Y. L. Yang, and A. J. Jacobson. Oxygen permeation studies of SrCo0 .sFeo.2 0 3 - delta. Solid State Ionics, 76(3-4):321 - 329, 1995. 153 [92] H. Kruidhof, H.J.M. Bouwmeester, R.H.E. v. Doorn, and A.J. Burggraaf. Influence of order-disorder transitions on oxygen permeability through selected nonstoichiometric perovskite-type oxides. Solid State Ionics, 63-65:816 - 822, 1993. [93] Jay Kniep and Y.S. Lin. Partial oxidation of methane and oxygen permeation in srcofeox membrane reactor with different catalysts. Industrial & Engineering Chemistry Research, 50(13):7941-7948, 2011. [94] Alexey Markov, Mikhail Patrakeev, Ilia Leonidov, and Victor Kozhevnikov. Reaction control and long-term stability of partial methane oxidation over an oxygen membrane. Journalof Solid State Electrochemistry, 15:253-257, 2011. 10.1007/s10008-010-1113-x. [95] Hui Dong, Zongping Shao, Guoxing Xiong, Jianhua Tong, Shishan Sheng, and Weishen Yang. Investigation on pom reaction in a new perovskite membrane reactor. Catalysis Today, 67(1-3):3 - 13, 2001. Catalysis in Membrane Reactors. [96] Zongping Shao, Hui Dong, Guoxing Xiong, You Cong, and Weishen Yang. Performance of a mixed-conducting ceramic membrane reactor with high oxygen permeability for methane conversion. Journal of Membrane Science, 183(2):181-192, 2001. [97] Jianhua Tong, Weishen Yang, Rui Cai, Baichun Zhu, and Liwu Lin. Novel and ideal zirconium-based dense membrane reactors for partial oxidation of methane to syngas. Catalysis Letters, 78:129-137, 2002. 10.1023/A:1014950027492. [98] A.A. Yaremchenko, A.A. Valente, V.V. Kharton, E.V. Tsipis, J.R. Frade, E.N. Naumovich, J. Rocha, and F.M.B. Marques. Oxidation of dry methane on the surface of oxygen ion-conducting membranes. Catalysis Letters, 91:169-174, 2003. 10. 1023/B:CATL.0000007150.63791.a2. [99] Stefan Diethelm, Joseph Sfeir, Frank Clemens, Jan Van herle, and Daniel Favrat. Planar and tubular perovskite-type membrane reactors for the partial oxidation of methane to syngas. Journal of Solid State Electrochemistry, 8:611-617, 2004. 10.1007/si0008-003-0489-2. [100] Haihui Wang, Rong Wang, David Tee Liang, and Weishen Yang. Experimental and modeling studies on Bao. 5 Sro.5 Coo.sFe0 .2O3. (BSCF) tubular membranes for air separation. Journal of Membrane Science, 243(1-2):405 - 415, 2004. [101] Shaomin Liu and George R. Gavalas. Preparation of oxygen ion conducting ceramic hollow-fiber membranes. Industrial & Engineering Chemistry Research, 44(20):76337637, 2005. [102] S. Diethelm, D. Bayraktar, T. Graule, P. Holtappels, and J. Van herle. Improved stability of Lao. 5Sro. 5 FeO 3 by Ta-doping for oxygen separation membrane application. Solid State Ionics, 180(11-13):857 - 860, 2009. Electroceramics XI - Ionic, Electronic and Mixed Conductors, Fuel Cells Symposium. 154 [103] Victor Kozhevnikov, Ilia Leonidov, Mikhail Patrakeev, Alexey Markov, and Yakov Blinovskov. Evaluation of La 0 .5 Sr0 .5 FeO 3 membrane reactors for partial oxidation of methane. Journal of Solid State Electrochemistry, 13:391-395, 2009. 10.1007/s10008008-0572-9. [104] Huixia Luo, Yanying Wei, Heqing Jiang, Wenhui Yuan, Yangxiao Lv, Jrgen Caro, and Haihui Wang. Performance of a ceramic membrane reactor with high oxygen flux Ta-containing perovskite for the partial oxidation of methane to syngas. Journal of Membrane Science, 350(1-2):154 - 160, 2010. [105] Tingfang Tian, Wendong Wang, Minchuan Zhan, and Chusheng Chen. Catalytic partial oxidation of methane over SrTiO 3 with oxygen-permeable membrane reactor. Catalysis Communications, 11(7):624 - 628, 2010. Catalytic partial oxidation of methane. [106] Yanying Wei, Hongfei Liu, Jian Xue, Zhong Li, and Haihui Wang. Preparation and oxygen permeation of U-shaped perovskite hollow-fiber membranes. AIChE Journal, 57(4):975-984, 2011. [107] Zhengliang Gong and Liang Hong. Integration of air separation and partial oxidation of methane in the Lao 4 Bao.6 Fe0 .8 Zn0 .2O3 membrane reactor. Journal of Membrane Science, 380(1-2):81 - 86, 2011. [108] Zigang Shen, Xiaorui Guo, Wuqing Zhang, and Lijun Pan. Effect of laser ablation onBao.5 Sr 0 .5 Coo. 8 Feo. 2 0 3 _ membrane for methane conversion. volume 279, pages 61 65, Hangzhou, China, 2011. [109] Xinzhi Chen, Liu Huang, Yanying Wei, and Haihui Wang. Tantalum stabilized SrCoO 3 perovskite membrane for oxygen separation. Journal of Membrane Science, 368(12):159 - 164, 2011. [110] Jianxin Yi, Jochen Brendt, Michael Schroeder, and Manfred Martin. Oxygen permeation and oxidation states of transition metals in (Fe, Nb)-doped BaCoO 3 perovskites. Journal of Membrane Science, 387-388(0):17 - 23, 2012. [111] P.N. Dyer, M.F. Carolan, D. Butt, R.H.E. Van Doorn, R.A. Cutler. Mixed conducting membranes for syngas production, US Patent 6492290, December 2002. [112] Liang Tan, Xuehong Gu, Li Yang, Wanqin Jin, Lixiong Zhang, and Nanping Xu. Influence of powder synthesis methods on microstructure and oxygen permeation performance of Ba 0 .5 Sro.5 Coo.sFeo. 20 3 _perovskite-type membranes. Journal of Membrane Science, 212(1-2):157 - 165, 2003. [113] A. Ellett, D. Schlehuber, L. Singheiser, and T. Markus. Permeation and stability investigation of Bao. 5Sro 5 Co 0 .sFeo.2 0 3 _ membranes for oxy-fuel processes. volume 30, pages 121 - 127, Daytona Beach, FL, United states, 2010. 155 [114] M. Pilar Lobera, Jos M. Serra, Sren P. Foghmoes, Martin Sgaard, and Andreas Kaiser. On the use of supported ceria membranes for oxyfuel process/syngas production. Journal of Membrane Science, 385-386(0):154 - 161, 2011. [115] Oliver Czuprat, Thomas Schiestel, Hartwig Voss, and Jurgen Caro. Oxidative coupling of methane in a bcfz perovskite hollow fiber membrane reactor. Industrial & Engineering Chemistry Research, 49(21):10230-10236, 2010. [116] F.T. Akin and Jerry Y.S. Lin. Oxygen permeation through oxygen ionic or mixedconducting ceramic membranes with chemical reactions. Journal of Membrane Science, 231(1-2):133 - 146, 2004. [117] Zebao Rui, Yongdan Li, and Y S Lin. Analysis of oxygen permeation through dense ceramic membranes with chemical reactions of finite rate. Chemical Engineering Science, 64(1):172-179, 2009. [118] Xiaoyao Tan, Yutie Liu, and K. Li. Mixed conducting ceramic hollow-fiber membranes for air separation. AIChE Journal, 51(7), 2005. [119] A. Thursfield and I. S. Metcalfe. Methane oxidation in a mixed ionic-electronic conducting ceramic hollow fibre reactor module. Journal of Solid State Electrochemistry, 10(8):604-16, 2006. [120] Alan Thursfield and Ian S. Metcalfe. Air separation using a catalytically modified mixed conducting ceramic hollow fibre membrane module. Journal of Membrane Science, 288(1-2):175 - 187, 2007. 156 Appendices 157 158 Appendix A Reactor within Enclosure 159 feed exhaust manifold gas mpling point feed inlet manifold mrem brahe pos tion enclosure heater opticaI access (n t in use in this pict ) Figure A-1: The ITM reactor within its enclosure in the laboratory. 160 Appendix B Methodology for Leak Quantification (under non-permeation conditions) The leak results described in section 3.1.3 are quantified (at temperatures lower than the onset of oxygen permeation) in molar flow rates to enable a comparison with the expected oxygen permeation flux across the membrane (1pmolcm-2s 1). With the values of Vsweepin (pure C0 2 ) and X0 2 ,8 ,0 t supplied by the CO 2 MFC and the GC respectively, the oxygen leak rate into the reactor can be quantified. There are two main considerations for this analysis. Firstly, we can assume that the concentrations of N2 (xN 2 ,s,ot) and CO 2 (xCo 2 ,S,out) in the sweep outlet stream are related to the concentration of 02 (Xo 2 ,s,out) in the sweep outlet stream thus: XN 2 0.79(B) 0.9 XO2,sout ,S,OUt (B.1) 1 Xco 2 ,s,oUt = 1 - (X0 2 ,S,OUt + XN 2,sOUt) = 1 - 02,SOUt Therefore, the average molar mass of the sweep side exit gas is 161 (B.2) M = (Xo2,S,out X Mo2 ) + (XN2 ,s,out X MN2) + (XCO2 ,S,OUt X MCO2) (B.3) Secondly, we can calculate the mass flow rate of exit sweep gas (rhsout) based on equa- tions 3.3 & 3.10. The mass fraction of 02 in the sweep side outlet (mfo2 ,S,out) is therefore M02 mfO2,8 ,out = XO2,s,out X M M With the above considerations, the oxygen leak into the reactor or across the membrane seal as detected on the sweep side outlet n02 ,leak (hO2,leak m4 "ut x2,Sout X mf = in mole-s- 1 ) can be quantified by: Onns~o M0 Or for direct analysis based on available data, niO ,leak 2 = x ro,~u ' 21 Xo 2,S, 0 ~t X (0.21 X02,8,out) v00 2 ,s,in 1 1 0 - 162 Pot PS t TetR (B.4) Appendix C Flow Procedures Legend V= Valve D = Flow Direction P = Vacuum Pump K= Key lock S = Switch DAQ = Data Acquisition System M = Mass Flow Controller MP = Mass Flow Controller Purge Figure C-1: Reactor flow diagram 163 Table C.1: Flow procedures Purpose Gas Used 1. 2. 1. 2. 3. 1. 2. Air Feed side inflow CO 2 Feed side purgea - Turn VI to the D2 direction Set M1 to desired flow rate Open V8 Set M2 to desired flow rate Turn V3 to the D4 direction Open V8 Turn on M2 3. Turn on MP2 CO 2 (h.pc) CO 2 CH 4 Sweep side purge Procedure CO2 (in) 2 CO 2 (h.p) Air d Reactor inflow (both sides) 4. Turn V3 to the D4 direction Open V2 1. Open V8 2. Set M2 to desired flow rate 3. Turn V3 to the D3 direction 1. Open V9 2. Set M3 to desired flow rate 3. Turn on K1, Si, and S2 1. Open V8 2. Turn on M2 3. Turn on MP2 4. Turn V3 to the D3 direction Open V5 1. Turn VI to the D2 direction 2. Set M1 to desired flow rate 3. Open V4 1. Open V8 2. Set M2 to desired flow rate 3. Turn V3 to the D3 direction 4. Continued on Next Page... 164 Open V4 Purpose Reactor purge (both sides) Enclosure purge Oxygen sensor calibration Table C.1 - Continued Procedure Gas Used 1. Turn VI to the D2 direction Air~ m2. Turn on M1 3. Turn on MP1 4. Open V4 1. Open V8 2. Turn on M2 3. Turn on MP2 CO 2 (m-p) 4. Turn V3 to the D3 direction 5. Open V4 Open V2 and V5 CO2 (h.p) u p1. Turn on S3 CO 2 and Air 2. Turn on P1 Open VI CO2 (0% 02) Open P-0 2 cylinder P-02 (10% 02) Turn V1 to the D1 direction Air (21% 02) aPurging involves using CO 2 to remove air from the reactor. bMass flow controller purge (total flow rate - 2000 sccm). cHigh purge (total flow rate > 2000 sccm). dAir flow in both sides (normal or purge) aids the cooling of the reactor by forced convection. 165