GAS TRANSMISSION THROUGH MICROPOROUS MEMBRANES Except where reference is made to the work of others, the work described in this dissertation is my own or was done in collaboration with my advisory committee. This dissertation does not include proprietary or classified information. Tacibaht Turel Certificate of Approval: ___________________________ Gisela Buschle-Diller Professor Polymer and Fiber Engineering ___________________________ Yasser Gowayed, Chair Professor Polymer and Fiber Engineering ___________________________ Peter Schwartz Professor Polymer and Fiber Engineering ___________________________ George T. Flowers Interim Dean Graduate School GAS TRANSMISSION THROUGH MICROPOROUS MEMBRANES Tacibaht Turel A Dissertation Submitted to the Graduate Faculty of Auburn University In Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy Auburn, Alabama December 19, 2008 iii GAS TRANSMISSION THROUGH MICROPOROUS MEMBRANES Tacibaht Turel Permission is granted to Auburn University to make copies of this dissertation at its discretion upon request of individuals or institutions and at their expense. The author reserves all publication rights. Signature of Author Date of Graduation iv VITA Tacibaht Turel, daughter of Nejla Turel and F. Tamer Turel, was born November 30, 1975, in Elazig, Turkey. She graduated from Elazig Anatolia High School in 1993. She graduated from Ege University, Izmir, Turkey, with a Bachelor of Science degree in Textile Engineering in June, 1997. After working in various apparel manufacturing companies in Turkey, she entered Graduate School, Auburn University, in May 2001. She completed a M.Sc. degree in Textile Engineering in December 2002. She completed a Ph.D. degree in Polymer and Fiber Engineering in December 2008. v DISSERTATION ABSTRACT GAS TRANSMISSION THROUGH MICROPOROUS MEMBRANES Tacibaht Turel Doctor of Philosophy, December 19, 2008 (M.Sc., Auburn University, December 19, 2002) (B.Sc., Ege University, Turkey, 1997) 170 Typed Pages Directed by Yasser Gowayed An ideal protective clothing material should be a good barrier against harmful gases or vapor while allowing moisture vapor and air passage through the material. In the study and design of barrier materials, one of the critical issues is to balance these requirements, which may sometimes be mutually exclusive. Therefore it is critical to understand the macroscopic and microscopic structure of the attack mechanisms as well as the barrier materials and the transport phenomena in such systems. In this study, air and gas transmission through barrier systems consisting of porous membranes was investigated experimentally and a molecular-level probabilistic model was constructed to evaluate the effect of various parameters on the gas flow. The effect of membrane parameters such as porosity, pore size distribution, thickness as well as gas parameters such as molecule diameters were examined at single layer as well as multiple layers. To understand the gas behavior for harmful chemicals and to ensure safety during vi experimental studies, mimics of such gases were obtained which were comparable to the actual gases in shape, molecular weight and other chemical properties. Air, ammonia and several mimic gases of harmful chemical agents were studied. Beta-pinene was used as a mimic of sarin and prenol was used as a mimic of nitrogen mustard. Gas transmission experiments were conducted on polyester, nylon and polypropylene membranes each of which had different porosity and pore size distributions. Experiments were done at different pressure values and a comparison was made between permeability testing machines based on volumetric and manometric principles as to their ability to accommodate high permeability membranes. Physical and chemical adsorption of such gases on porous membranes was also investigated after the addition of active elements on the membrane surfaces which can interact with the gas molecules. An experimental setup was developed to measure concentration changes upon passage of the gas through active and non-active membranes by which way the effect of active elements on gas transmission rates could be examined. Probabilistic and mathematical models were proposed to predict gas transmission rates of such membranes and gas permeation experiments were simulated utilizing the proposed models. Modeling results were compared to experimental data and to macro- level gas transmission models available in literature. The results were found to be in acceptable correlation with experiments for membranes with a well-known morphology. The pressure of the system, porosity and adsorption capacity of the materials, as well as size of the gas molecules and their interaction energies between the surfaces were found to be important parameters which affect permeability. vii ACKNOWLEDGEMENTS I would like to thank Dr. Yasser Gowayed for his invaluable guidance, understanding, support, wisdom, patience and faith in me during the last five years. I would also like to thank Dr. Ebraheem Shady, who has been a very valuable source of knowledge, for all his support and help during this study. I am very grateful to Dr. Gisela Buschle-Diller for her gracious support in many ways during the course of this study. I am very thankful to Dr. Tamer Awad for his help during the final part of my work. I would also like to thank all my professors and fellow graduate students at Auburn University for their help, patience and support at different stages of my studies as well as my family and friends without whose support I could not have completed this work. I would like to thank National Textile Center and Department of Homeland Security for providing financial support. viii Style manual or journal used Guide to Preparation and Submission of Theses and Dissertations, Textile Research Journal _____________________________________________________________________ Computer software used Matlab, Gaussian, Pymol, Griffin _____________________________________________________________________ ix TABLE OF CONTENTS LIST OF FIGURES....................................................................................................xiii LIST OF TABLES ...................................................................................................xvii CHAPTER I..................................................................................................................1 LITERATURE REVIEW..............................................................................................1 1.1. Introduction...................................................................................................1 1.1.1. Purpose and objectives..........................................................................2 1.2. Membranes as Barrier Materials...................................................................3 1.3. Definitions of Important Parameters in Transport Through Porous Media..5 1.4. Types of Transport Through Membranes......................................................7 1.5. Principles of Gas Transport through Membranes.......................................10 1.5.1. Molecule-wall collisions.....................................................................14 1.5.2. Molecular collisions and viscous vs. Knudsen flow...........................15 1.5.3. Diffusion in Knudsen regime..............................................................16 1.5.4. Knudsen ? Poiseulle transition model.................................................18 1.6. Flow rate of gases through larger openings and tubes................................20 1.7. Chemical and Physical Adsorption of Gases..............................................22 1.8. Evaporation of Liquids................................................................................26 1.9. Microscale Models for Permeation of Gases through Protective Materials28 1.9.1. Mathematical models for fibrous materials.........................................28 1.9.2. Mathematical models for membranes.................................................30 x 1.9.3. Molecular simulations and modeling..................................................33 1.9.4. Adsorption studies...............................................................................34 1.10. References...................................................................................................36 CHAPTER II...............................................................................................................41 GAS TRANSMISSION RATE THROUGH MICROPOROUS MEMBRANES......41 2.1. Introduction.................................................................................................41 2.2. Experimental Work.....................................................................................44 2.2.1. Microporous membranes.....................................................................44 2.2.2. Evaluation of gas transmission rate.....................................................47 2.2.3. Evaluation of porosity and pore diameter...........................................49 2.3. Probabilistic Model.....................................................................................49 2.4. Results and Discussion................................................................................51 2.5. Conclusions.................................................................................................63 2.6. References...................................................................................................64 CHAPTER III..............................................................................................................67 TRANSPORT BEHAVIOR OF MIMICS of WARFARE AGENTS THROUGH POROUS MEMBRANES...........................................................................................67 3.1. Introduction.................................................................................................67 3.2. Materials......................................................................................................69 3.3. Experimental Work.....................................................................................75 3.4. Evaporation Model......................................................................................78 3.5. Results and Discussion................................................................................83 3.5.1. Experimental results............................................................................83 xi 3.5.2. Comparison of experimental data to modeling results........................86 3.6. Conclusions.................................................................................................90 3.7. References...................................................................................................91 CHAPTER IV.............................................................................................................93 CHEMICAL ADSORPTION BEHAVIOR OF MICROPOROUS MEMBRANES WITH ACTIVE ELEMENTS.....................................................................................93 4.1. Introduction.................................................................................................94 4.2. Adsorption Phenomena...............................................................................95 4.3. Membrane and Gas Properties....................................................................97 4.4. Modeling of Chemical and Physical Adsorption........................................98 4.4.1. Energy calculations.............................................................................98 4.4.2. Probabilistic chemi-sorption model..................................................102 4.5. Experimental Setup...................................................................................105 4.6. Results and Discussion..............................................................................108 4.6.1. Calibration Measurements for Concentration...........................................108 4.6.2. Concentration experiments with Gas Chromatography....................110 4.6.3. Comparison of permeation rates by the chemi-sorption model........117 4.7. Conclusions...............................................................................................119 4.8. References.................................................................................................120 CHAPTER V.............................................................................................................122 CONCLUSIONS AND FUTURE WORK...............................................................122 5.1. Conclusions...............................................................................................122 5.2. Future Work and Improvements...............................................................124 xii APPENDIX A...........................................................................................................126 TESTING INSTRUMENTS FOR MATERIAL PROPERTIES..............................126 A.1. LYSSY Air Permeability Tester...............................................................126 A.2. Gurley Permeability Tester.......................................................................127 A.3. Scanning Electron Microscope (SEM)......................................................128 A.4. Capillary Flow Porometer.........................................................................130 APPENDIX B...........................................................................................................131 ADDITIONAL EXPERIMENTS AND RESULTS.................................................131 B.1. Permeability Results with Treated PET and Nylon Membranes...............131 B.2. Permeation Tests with Gurley...................................................................133 B.3. Comparison of Results with Gurley and LYSSY Permeability Testers....134 APPENDIX C...........................................................................................................136 SOURCE CODES.....................................................................................................136 C.1. Matlab Code for Permeation of Gases through Porous Membranes.........136 C.2. Matlab Code for Permeation of Mimics through Porous Membranes......139 C.3. Matlab Code for Permeation of Gases through Membranes with Active Elements................................................................................................................143 C.4. Subroutines................................................................................................148 C.4.1. Chemisorption...................................................................................148 C.4.2. Porecount...........................................................................................149 C.4.3. Porosity..............................................................................................150 C.4.4. Knudsen.............................................................................................150 C.4.5. Trajectory..........................................................................................151 xiii LIST OF FIGURES Figure 1.1. Flow mechanisms (a) viscous flow, (b) Knudsen flow, (c) surface flow (molecular sieving), (d) solution-diffusion.................................................................10 Figure 1.2. Knudsen number vs. flow distribution [4]................................................12 Figure 1.3. Required energy levels for adsorption [redrawn from Ref. 3].................24 Figure 1.4. Effect of Kinetic Energy vs. Surface Potential Energy on the molecules inside the pore.............................................................................................................26 Figure 2.1. SEM images of microporous membranes: (a) PET (b) B130, and (c) nylon .....................................................................................................................................45 Figure 2.2. Cross-section SEM images of double layer PET membranes with incorporated microparticles.........................................................................................46 Figure 2.3. LYSSY L100-5000, air permeability measurement chambers [7]...........47 Figure 2.4. CSI-135, view of the permeation cell [8].................................................47 Figure 2.5. The sketch of a possible molecular trajectory..........................................50 Figure 2.6. Air transmission rate of microporous membranes....................................52 Figure 2.7. Transmission rate of ammonia through microporous membranes............53 Figure 2.8. Comparison of air and ammonia transmission rates through PET...........54 Figure 2.9. Air transmission rate through double layered PET membranes...............55 Figure 2.10. Interaction energy profiles for air and ammonia with PET...................57 xiv Figure 2.11. Interaction energy profiles for air and ammonia molecules and PET surface covered with a thin layer of epoxy.................................................................58 Figure 2.12. Comparison of experimental data and numerical models for the permeation of air through PET membrane..................................................................59 Figure 2.13. Comparison of experimental data and numerical models for the permeation of air through nylon membrane................................................................59 Figure 2.14. Comparison of experimental data and numerical models for the permeation of air through B130 membrane................................................................60 Figure 2.15. Comparison of experimental data for permeation of ammonia and the probabilistic model for uncoated PET microporous membrane, with and without the interaction energy between molecules and the surface...............................................62 Figure 2.16. Comparison of experimental results and model for double layered PET .....................................................................................................................................63 Figure 3.1. Sarin Gas...................................................................................................71 Figure 3.2. Beta-pinene (mimic for sarin)...................................................................71 Figure 3.3. Nitrogen Mustard......................................................................................72 Figure 3.4. Prenol (3-methyl 2-buten 1-ol) (mimic for mustard)................................72 Figure 3.5. Vapor pressure of ?-pinene [7, 10]...........................................................74 Figure 3.6. Vapor pressure of prenol [7, 10]...............................................................74 Figure 3.7. SEM images of microporous membranes: (a) PET (b) nylon..................75 Figure 3.8. Schematic of bottom chamber (left) and its image (right) with thermocouples wires....................................................................................................76 Figure 3.9. Modified setup for measurement of permeation of mimics......................77 xv Figure 3.10. Slug movement versus Pressure relationship.........................................81 Figure 3.11. Pressure versus time during slug movement due to evaporation of mimics.........................................................................................................................81 Figure 3.12. GTR vs. Temperature for ?-pinene........................................................83 Figure 3.13. GTR vs. Temperature for Prenol............................................................84 Figure 3.14. Comparison of beta-pinene and prenol through PET.............................85 Figure 3.15. GTR versus temperature for evaporation of beta-pinene through PET membrane....................................................................................................................87 Figure 3.16. GTR versus temperature for evaporation of beta-pinene through nylon membrane....................................................................................................................87 Figure 3.17. GTR versus temperature for evaporation of prenol through PET membrane....................................................................................................................88 Figure 3.18. GTR versus temperature for evaporation of prenol with nylon..............89 Figure 4.1. Chemical structure of XL-I-77.................................................................97 Figure 4.2. Adsorption of NH3 on modified PET with 5-azido-3-oxapentanesulfonate .....................................................................................................................................98 Figure 4.3. Interaction energy between the gases and the surface covered with PGMA .....................................................................................................................................99 Figure 4.4. Interaction energy between ammonia and the surface with active element (5-azido-3-oxapentanesulfonate)...............................................................................100 Figure 4.5. Interaction energy between mustard gas and prenol and the surface with active element (XL-I-177).........................................................................................100 Figure 4.6. Interaction energy for ammonia..............................................................101 xvi Figure 4.7. Determination of chemisorptions field...................................................103 Figure 4.8. Ammonia reaction with benzoyl chloride...............................................106 Figure 4.9. Schematic view of adsorption testing setup............................................107 Figure 4.11. (a)The actual mass spectrum of benzamide from calibration experiments, (b) expected mass spectrum for benzamide..............................................................108 Figure 4.12. Calibration curve for benzamide concentration measurements............109 Figure 4.13. Chromatography results for pure benzoyl chloride (a) and pure benzamide (b) in methanol........................................................................................111 Figure 4.14. Gas Chromatography peaks for benzamide inside the reaction solution with 10 ?l BOC and 238.6 ml/min ammonia passing through untreated PET.........113 Figure 4.15. Gas Chromatography and mass spectrum for benzamide inside the reaction solution with 5 ?l BOC and 238.6 ml/min ammonia passing through untreated PET............................................................................................................114 Figure 4.16. Gas Chromatography and mass spectrum for benzamide inside the reaction solution with 5 ?l BOC and 238.6 ml/min ammonia passing through active PET............................................................................................................................115 Figure 4.17. Comparison of models for air and ammonia with and without adsorption on PET.......................................................................................................................118 Figure A.1. The LYSSY 100-5000 Permeability Tester and the humidity control..127 Figure A.2. Rotary pump (left) and heated/refrigerated circulator (right)................127 Figure A.3. Model 58-03 Gurley Densometer..........................................................128 Figure A.4. (a) MF715, (b) W-gray, (c) Aptra Classic, (d) B130, (e) W-wide, (f) UV8 ...................................................................................................................................129 xvii Figure B.1. Gas transmission rate of PET with and without treatment at two different pressures....................................................................................................................132 Figure B.2. Gas transmission rate of nylon with and without treatment at different temperatures..............................................................................................................132 Figure B.3. Gurley experiments for PET, nylon and B130 membranes...................133 Figure B.4. Permeance of PET for single, double and three layers..........................134 Figure B.5. LYSSY and Gurley comparison.............................................................135 xviii LIST OF TABLES Table 2.1. Geometric properties of microporous membranes...............................44 Table 3.1. Molecular diameter of air components, toxic agents and their mimics73 Table 3.2. Geometric properties of microporous membranes...............................75 Table 4.1. Concentration differences between untreated and active PET microporous membranes???..............................................................................................116 1 CHAPTER I LITERATURE REVIEW 1.1. Introduction Chemical Protective Clothing (CPC) is used to isolate the skin from direct contact with potentially harmful chemicals and minimize the risk of injury. In warfare, protective clothing must act as a barrier to chemical and biological warfare agents and prevent the hazardous chemicals or bacteria from reaching to the skin. The hazardous chemicals may be in gas, liquid or solid form and the protective materials must be designed based on the type of chemicals to which they will most likely be exposed. The performance of CPC mainly depends on the resistance of fabric to chemical permeation; however, there are other factors that can be very critical for specific purposes of clothing, such as a battle dress or a hazardous material suit which may require different protection levels. Studies show that protective clothing materials which have good barrier performances hinder the performance of the wearer, due to heavy weight, heat stress and limited breathability [15, 32, 37, 43]. On the other hand, increased breathability may reduce the protection level. Therefore, it is very critical to analyze these requirements, which appear to be mutually-exclusive, to determine the most appropriate materials and design for chemical protective clothing. To understand how an individual can be protected from the harmful effects of the chemicals, it is necessary to understand the 2 macroscopic and microscopic structure of the attack mechanisms as well as the barrier materials. Comfort is directly related to stiffness, heaviness, and air, heat and moisture transfer properties of the fabric. Barrier and comfort parameters must be understood at the fiber, yarn, fabric as well as at the garment level [47]. For instance at the fabric level, important parameters for consideration, beside barrier effectiveness, include comfort limit, resistance to abrasion and contamination of the material [5, 56, 9]. Barrier effectiveness of a textile material is closely related to the fabric structure, its porosity, pore size and thickness [35]. There is a diverse range of technical issues involved in developing comfortable barriers. Barker et al. [5] defined the comfort limit as ?the range of body activity levels within which an individual wearing a clothing system is predicted to be thermophysiologically comfortable?. An ideal protective clothing material should not allow passage of harmful gases or vapors, while allowing moisture vapor to move out at any pressure. Shortly, an ideal protective material should have high moisture vapor permeability and low gas permeability [46]. 1.1.1. Purpose and objectives The main question in the study of protective textiles against hazardous chemicals is how to manufacture materials which have effective barrier properties as well as a sufficient comfort level. Therefore, the primary objective of this study is to provide a better understanding of the barrier and comfort relationship in protective materials. To do that, a multi-layer structure consisting of fabric and microporous membrane layers is proposed for consideration and air and gas transmission through these materials is studied 3 experimentally. Since the barrier ability of the fabric is negligibly small compared to that of the membrane layers, the gas transmission studies are focused on the microporous membranes. Therefore, the objectives of this study can be listed as follows: ? To study gas transmission as well as transmission of mimics of hazardous chemical gases through microporous membranes experimentally ? To study microporous membranes with active elements for chemical blockage of the gases and its effect on transmission rates experimentally ? To construct a model which can evaluate the transport phenomenon through both virgin and activated membranes at single and multiple layers for gases and mimics ? To compare experimental findings with the model and finally, ? To provide a guide for developing rules and standards for the design of breathable and effective chemical protective clothing materials. 1.2. Membranes as Barrier Materials Chemicals may affect a protective material in three different processes; degradation, penetration and permeation. Degredation is the change or deterioration in the material properties caused by surface contact by the chemical. Penetration is usually considered the flow of a liquid or gas on a non-molecular level through the seams, holes or larger pores, while permeation is the flow of fluids through the material on a molecular level. Permeation can occur in the form of adsorption, diffusion or desorption [54]. Polymeric membranes and films are commonly used for protection against liquid and gas penetration. Those membranes are laminated or bonded to other textiles ?typically 4 non-wovens- to provide strength and durability. Membranes or films can be porous (permeable) or monolithic (impermeable). A porous film permits movement of vapor via microscopic channels across the material. In a monolithic film, on the other hand, vapor transmission occurs mainly by molecular diffusion. A more detailed categorization of barrier materials can be listed as follows [46, 47]: 1. Permeable: most fabrics and materials containing activated carbon-layers to adsorb toxic vapors. 2. Semi-permeable: membrane structures with low air permeability and high water- vapor permeability. 3. Selectively-permeable: structures that allow permeation of smaller molecules and water-vapor, while preventing the passage of larger organic molecules. 4. Impermeable: Hazardous Material response suits which are completely impermeable such as those made of Tyvek-reinforced polypropylene film laminates or aluminized Kevlar fabrics. Membranes can also be categorized based on different criteria in different studies. The most common way is to group them depending on the structure of the pores. Membranes with a pore size less than 0.7 nm are considered as fully dense membranes. Membranes with pore sizes greater than 50 nm are considered macroporous membranes [41]. In some studies membranes with pores between 200 and 3000 nm are considered microporous membranes [13]. Although most studies have their own categorization of micro and macroporous membranes, International Union of Pure and Applied Chemistry divides these membranes into two further categories and defines membranes with pore 5 sizes between 2 nm and 50 nm as mesoporous and with pore sizes not exceeding 2 nm as microporous [7, 52]. Although most membranes used in this study can be considered macroporous based on this categorization, they were mentioned as microporous membranes throughout this dissertation as they were understood in many of the studies as well as by the manufacturer. Since they are usually layered with fabrics which have very large openings in comparison, we believe it is appropriate to consider fabric openings as macro pores, while membrane pores are comparatively considered micro. 1.3. Definitions of Important Parameters in Transport Through Porous Media Many researchers have studied aspects of air or gas permeation through protective materials, however the permeation through micro pores in the case of a multi-layer assembly has not been sufficiently understood. There are many factors which affect the transport of gas or vapor through porous materials. In impermeable membranes, the type of polymer and membrane thickness are the most important factors in determining the transmission rates, while for microporous membranes, gas and vapor transmission is closely related to porosity, the overall thickness and the mean pore diameter [36]. Similarly, in textile fabrics, the role of porosity and pore geometry is significant in determining the transport behavior [35]. Porosity is defined as ?the ratio of the volume of air or void contained within the boundaries of the material to the total volume (solid material plus air or void) expressed as a percentage [1]?. It has also been defined as the ratio of the geometrical area of the opening on the surface to the total area of the material [24]. In fabrics, porosity can be calculated using an optical image analysis system, which transmits light through the 6 fabric and calculates the open area of the pores [35]. In this case, it is possible to talk about an optical porosity. Shelekhin et al. [48] considered the abovementioned definitions of porosity as total porosity and introduced the term throughout porosity for membranes which contain inlet and outlet boundaries, i.e. isolated and dead-end pores, throughout their thickness. When the length of the diffusion path is higher than the length of a straight capillary, a tortuosity factor is considered. Tortuosity (?) is an empirical coefficient which is expected to be as 1>1 or viscous if Kn<0.01 (Figure 2). The flow within the intermediate region of the Knudsen number interval is called the transition flow. The regime between Knudsen and viscous is often called the slip flow. In all three of these cases, the molecule diameter is 12 considered to be very small compared to the pore diameter. As pore diameter gets smaller and comparable in size to the molecule diameter, it becomes a configurational diffusion (surface flow). Although most of the gas flow in porous media lay in the transition region, there is a lack of analytical or numerical models to describe the transmission rate within this region. Kn 10 0.1 0.01 0 Figure 1.2. Knudsen number vs. flow distribution [4] Although the true nature of flow cannot be defined by a mere evaluation of mean free path and Knudsen number, this information can be useful in understanding the general flow mechanism and enable a better estimation of the flow characteristics. A summary of the evaluation of types of fluxes based on various parameters such as the ratio of the molecular size to the pore diameter or Knudsen number is shown below [12]: m? ? <<1 and pr ? <<1 barb2right Viscous Regime m? ? >>1 and pr ? <<1 barb2right Knudsen Regime m? ? ~1 and pr ? <<1 barb2right Slip Regime Transition Free-molecular Continuum Slip 13 m? ? >>1 and p m r ? >>1 barb2right Pseudo-Knudsen (Transitional) Regime m? ? <<1 and pr ? <<1 barb2right Molecular m? ? <<1 and pr ? <<1 barb2right Configurational When the behavior of a gas is analyzed, it is done by looking at a large population (macroscopically) instead of individual molecules (microscopically), since the movement of individual molecules is chaotic, while the whole system can be examined in the light of gas laws. The macroscopic system usually behaves based on concentration, temperature, pressure, etc, although individual molecule movements are probabilistic. This macroscopic analysis is called the continuum hypothesis [12]. However in most cases, the changes in the individual molecules are so high that it affects the population as a whole and the continuum hypothesis become inapplicable. In addition, in continuum regimes the interaction between gas molecules and surface of porous media is not taken into consideration. Therefore, it is more appropriate to examine movement of gas molecules microscopically, in a probabilistic manner, rather than at a macroscopic level. Velocity and diffusion of gases take place through collisions between molecules. As the number of molecules in a unit volume increases, the frequency of the collisions increases which increases the pressure of the system. If a gas mixture consists of molecules of type which is not confined with walls, the diffusion takes place through molecule-to-molecule collisions. This is called molecular diffusion. Since molecule-wall 14 collisions do not exist in this case, momentum is transferred from one molecule to the other and the total momentum is conserved for the whole system [12]. 1.5.1. Molecule-wall collisions Gas transport in porous media occurs usually in systems that are confined with walls, which introduces molecule-wall collisions. These systems are more complex since the movement of the molecule depends on several factors such as the roughness of the surface and the interaction of the surface with the molecule, etc. Three different situations may occur: ? Elastic Collisions: If the wall has a smooth surface and the collision is elastic, the momentum is conserved similar to a system without walls, with the exception of a reverse velocity component in this case. In elastic collisions, the angle of the molecule hitting the wall is the same as the angle of the reflection from the wall. Although one component of velocity is reversed in direction, the molecule still moves in the direction of the whole system, with the same magnitude of velocity. ? Diffusive Collisions: If the wall has a rough surface, the system will loose momentum after collisions and the movement of individual molecules will be chaotic. Neither the reflection angle nor the magnitude of the velocity will remain the same. ? Adsorption by the Wall: In this case, the reflection of the molecule from the wall depends on the attractive force of the wall. This collision often causes surface diffusion, since the molecule is forced to migrate over the surface. 15 Due to the chaotic nature of their reflection, molecules that collide on adsorptive surfaces are not distinguishable from the ones that collide on rough surfaces at a macroscopic scale. Therefore, it is possible to categorize collisions only in two categories; as either elastic or non-elastic. Since every molecule has some degree of interaction with the surfaces, most collisions with surfaces are usually non-elastic at some degree. In the absence of adsorption conditions, some studies have assumed all collisions to be elastic for simplicity. 1.5.2. Molecular collisions and viscous vs. Knudsen flow When all the molecules that collide with the wall rebound and collide with other molecules, this flow is described as a viscous flow. Viscous flow may have a viscous, or turbulent regime, depending on how much of the momentum change can be transferred to the entire fluid. Although molecules have a chaotic nature in their collisions, in viscous flow the bulk of the fluid has a direction and an overall velocity [12]. Flow through highly porous media is often viscous. In Knudsen flow, the number of molecule-wall collisions is much higher than that in viscous flow that the molecule-molecule collisions become negligible. Therefore, the flow does not have an overall direction. The pressure only affects the concentration of the species; therefore Knudsen flux is directly proportional to the pressure of the system. When there are more than one species in the fluid, they may not always diffuse as a complete mixture. When the mixture does not have an integration of the species throughout the process, the flow is called a segregative flow. In Knudsen flow with more than one species, each species will have their own flux, independent of each other, 16 whereas in molecular diffusion, different species are interdependent in their behavior. In viscous flow, the flux is not segregative for different species. Since, Knudsen flow is based on molecule-wall collisions, after the molecule hits and rebounds from the wall surface, it can be assumed that the molecule will not collide with other molecules, therefore has no resistance to diffusion. In this regime, the velocity of the molecule is not zero at the surface and can be assumed to be equal at any distance from the wall compared to a viscous flow. In other words, it is independent of the distance from the wall. However in viscous flow the velocity can be assumed as zero at the surface. It gradually reaches a steady state as it gets away from the wall. For viscous flow, when the pressure is uniform, the total momentum loss as well as the flux is zero. For Knudsen flow, since the only collision is between the molecule and the wall, velocity distribution is the same for the whole system, regardless of distance from the wall. For viscous flow, velocity distribution is different from the bulk of the fluid near the wall [12]. To summarize, Knudsen flow depends on the mean distance between molecule-wall collisions (mean free path). Viscous flow depends on the mean hydraulic radius, which is the ratio of the cross-sectional area of the flow to the affected perimeter of the pore [45]. 1.5.3. Diffusion in Knudsen regime Gas flow changes depending on several parameters, such as temperature, pressure and concentration. For an ideal gas in laminar flow in a circular tube, with the assumption that all collisions are diffusive and not elastic, i.e. in a Knudsen diffusion, the mass rate of flow can be calculated using the Haagen-Posieulle equation as follows [8]: 17 ?????? ?= L ppaTm L03)34(2 ???? Where, m=molecular mass, ?=Boltzmann constant, a=tube radius and L=tube length. The molar flux can be calculated as: dz dc M RTa dz dp RTM aN A A A A A ?? 23 8 2 1 3 8 ?=?= This equation can be re-written as follows: dz dcDN A AKA ?= Where, AKD =Knudsen Diffusivity. However, to obtain the molar flux based on the total cross section of the porous medium, we have to include the porosity and tortuosity of the material. In that case, the molar flux becomes: dz dcDN Aeff AKA ?= Where, AKeffAK DD )/( ??= , ?=porosity and ?=tortuosity. Additionally, using Fick?s second law, the total volume of a gas that permeates through a porous medium Q in time t can be calculated as follows: 18 dttCDAQ t hxt ? ?? ?? ? ? ? ??= = 0 Where, h=membrane thickness and A=surface area. 1.5.4. Knudsen ? Poiseulle transition model In most microporous membranes, a pure Knudsen or a pure viscous flow is hard to achieve. As mentioned before, if the diameter of the pores is too large for a pure Knudsen diffusion, compared to the molecule diameter, the flow is considered in a transition regime between Knudsen and viscous [45]. Depending on the size of the pores which the individual molecules enter into; both a Knudsen-style and a viscous-style behavior may exist in the same material. For instance, at a certain temperature and pressure, the flow of nitrogen with a mean free path of 0.13?m through pores of 0.1-0.5 ?m, is not purely Knudsen or purely viscous, but in a transition regime. Schofield et al. [45] calculated the flux in the Knudsen region for an isotropic, microporous membrane as: ???? ? PKP RT M M RTrJ KK ??? ?? ? ?? ?= 2/18 3 2 Where, ? = membrane thickness, r = pore radius, ? = tortuosity, K=permeability of the material, _8 uMRT =? =mean molecular speed. Viscous flow is calculated as follows: ???? ? PKP RT MPrJ PP ???= 1 8 1 2 19 Where, ?=gas viscosity. In this equation, mean molecular velocity (?) is calculated by utilizing the Maxwell speed distribution and molecular mass. For a gas at equilibrium, the molecular velocities are random in nature and have an average magnitude which is calculated as [8]: MRTu ?8 _ = Most available models to describe the transition regime flux are in general non- practical due to their complexity and the required detailed knowledge of membrane morphology. Schofield et al. [45] developed a model which can can be used to correlate the flux to the membrane parameters with a semi-empirical relationship. In this model, membrane permeability is described with the equation: baK ? ? = Where, a=membrane constant, b=Knudsen/Poiseulle constant (b=0 for Knudsen, b=1 for Poiseulle) and ?=dimensionless pressure ( refP P ). This gives the flux as: PaJ b?= ? The effect of pressure in the flux changes based on the value of constant b. As b goes to 0, the flux in transition regime gets closer to a Knudsen-type behavior. The Poiseuille 20 flux can be re-written as follows, with the incorporation of mean free path which replaces the viscosity of the gas in the equation: ??? ? ? ? P RT M M RTrrJ P ? ?? ? ?? ?= 2/18 32 If the Knudsen and Poiseuille equations are linearly added together, the following equation is obtained: K=M ? (A+B/?) Where, A=2r?/3?RT, B=?r2?/32?RT, and ?=(8RT/?M)1/2. A and B are membrane constants for a given temperature. When this equation is refined, the initial equation baK ? ? = can be derived. Where ? is a property of the membrane and is equal to ?K at a reference pressure Pref, the term b is a measure of the operating regime and shows the contribution of Poiseuille flow in the transition. Both terms (a and b) are independent of each other unlike most other models where the constants have a dependency on each other. 1.6. Flow rate of gases through larger openings and tubes Isothermal flow of gases and liquids is derived from the equation of motion below by making certain assumptions and simplifications. ].[v gpDtD ??? +????= When equation of motion is simplified for constant ? (density) and ? (viscosity), the Navier-Stokes equation can be obtained as follows: 21 vv or vv 22 ?+??=+?+??= ????? PDtDgpDtD Where, DtD is the substantial time derivative, P is the modified pressure (P=p+?gh) and v is the velocity. The modified pressure is defined where h is the elevation in the gravitational field and gh is the gravitational potential energy per unit mass. Creeping or Stokes flow equation can be obtained by neglecting the acceleration terms in the Navier- Stokes equation, i.e. by assuming DtD? =0. When the viscosity of the fluids are relatively low, the fluids can be assumed to be inviscid fluids and viscous forces can also be neglected, which results in the Euler equation. When the flow is assumed to be a steady flow, then the Bernoulli equation can be derived from the Euler equation for steady and irrotational flows as follows [8, 16]: constant)(21 22 =++ Pyx ??? At low velocities, this equation can simply be written as: P?=221?? Where, ?P=the difference between the total pressure and the static pressure. If the pressure is known, this equation can be used to calculate volumetric flow rate (w) as w=?vS, where S=cross-sectional area. When the flow passes through an orifice or different cross-sectional areas, the cross-section will play an important role in the velocity profiles and therefore affect the rate of flow. In this case, the velocity change should be calculated which will be proportional to the ratio of the cross-sectional area [8]. 22 1.7. Chemical and Physical Adsorption of Gases Adsorption is usually defined as the attachment of gases or liquids onto a solid surface. Unlike absorption, adsorption occurs on the surface and does not cause the gases or liquids to dissolve in another liquid or solid. A solid that has an interaction with the gas molecules is called the adsorbent. The gas or vapor molecule is called the adsorptive prior to attaching to the surface and adsorbate upon adsorption onto the surface [16]. The interaction of gas molecules with a solid is a complex phenomenon due to the nature of the molecules and their variability in size, structure and electrical properties [29]. Adsorption levels of molecules onto a solid surface depend strictly on their energy loss close to that surface. When there is no energy loss, the molecules are considered to have an elastic collision with the surface without any bonding. If there is small energy loss, the collision is considered inelastic. When sufficient energy loss occurs, adsorption takes place. Depending on the amount of energy loss required, adsorption may occur in two different ways: a) Physical adsorption: In this process, a weak, physical interaction occurs between the gas and the solid and the electronic structure of the molecule is not affected by the adsorption process. The molecules are bound to the surface with van der Waals forces which require lower adsorption energies, i.e. less energy loss and can be easily reversible. It can also occur on multiple layers. b) Chemical adsorption: This process is usually a result of strong covalent or ionic bonds which are not easily reversible. They require higher adsorption energies and usually only happen on adsorptive sites and as a monolayer [3, 57, 34]. Chemisorption can also be associative or dissociative. If adsorption requires dissociation of a molecule 23 into its atoms before adsorption, it is called a dissociative adsorption. If the adsorbate is a single-atom molecule or dissociation is not required for adsorption, it is an associative adsorption [3]. As mentioned above, the adsorption process is a result of energy loss of the molecule close to the surface. This energy loss occurs based on the kinetic energy of the molecule as well as the attraction between the molecule and the surface which is related to the surface potential energy. If the attraction between the molecule and the surface is high, the molecule is more likely to experience energy loss, i.e. lose its kinetic energy and become adsorbed by the surface. The two important energy parameters are defined as translational kinetic energy of the molecule (Ek) and the potential/interaction/attraction energy of the surface versus gas (Ep or Ea). Energy loss of the molecules is related to the distance of the molecule from the solid [40]. For instance, at a distance where kinetic energy of the molecule is lower than the interaction energy on the surface, an adsorption is expected to take place. If the distance between the molecule and the surface is close enough this will be in the form of a chemical adsorption since chemical adsorption usually requires a higher energy loss, i.e. a closer distance from the wall. Physical adsorption can occur at further distances from the wall with less energy loss [3, 57]. When there is strong attraction between the gas and the solid, chemical adsorption has priority over physical adsorption [34]. 24 Figure 1.3. Required energy levels for adsorption [redrawn from Ref. 3] A schematic view of physisorption, chemisorption and the required attraction energies relative to distance from the surface is shown in Figure 3. The rate of adsorption depends on several parameters such as pressure, temperature, surface area, surface energy distribution and porosity of the solid material. The relationship between the number of molecules that are taken up by the surface and the pressure at constant temperature is referred to as the adsorption isotherm [57]. In the evaluation of adsorption behavior of a system reaching equilibrium, Langmuir adsorption isotherm is used most commonly with several assumptions and approximations. For chemisorption, Langmuir isotherm equation assumes that: a) There are only a fixed number of sites available for adsorption each of which can only be occupied by one molecule, b) At constant temperature, there is a dynamic equilibrium between the adsorbed layer and the gas at pressure P 25 c) The adsorbate molecules are constantly colliding with the surface. If they collide with an available adsorption site, they are adsorbed. If they collide with a filled adsorption site, they bounce back into the gas phase without sticking on the surface. Rate of adsorption can be calculated as follows based on Langmuir isotherm for chemical adsorption: Rate of adsorption=kaP(1-?) Where, P is the pressure, ka is the rate constant for adsorption and ? is the fractional single layer coverage of the surface by the molecules. Similarly, rate of desorption can be calculated as follows: Rate of desorption=kd? Where, kd is the rate constant for desorption. At equilibrium, the rate of adsorption and desorption will be equal: kd?= kaP(1-?). The fraction of coverage on the surface can be written as follows: KP KP N Nads +== 1? Where, K= d a k k , N ads=number of adsorbed molecules and N=available active sites. Where all sites are filled ?=1. At zero pressure, adsorption may be too slow and equilibrium may not occur easily, while at higher pressures, the equilibrium can be reached at a faster rate [3, 57]. Shindo et al. [50] considered the interaction energy between gas molecules and the pore surface by comparing the interaction energy to the kinetic energy of the gas molecule. They calculated that the gas molecule can either elastically collide with the 26 pore surface or be trapped in the potential field and hop on the surface. If the kinetic energy (Ek) of the molecule is greater than the interaction energy or potential energy of the pore surface (Es), the molecule will be able to escape from the surface potential field and rebound elastically without losing energy (Figure 4). In this case, the molecule will be able to proceed inside the pore in the gas phase. Alternatively, if the interaction energy between the surface and the gas molecule is greater than the kinetic energy of the molecule, then the molecule will be trapped in the potential energy field and physisorbed on the surface, in which case the movement of the molecule will be in the form of surface flow along the pore wall. This means that the molecule will lose some of its kinetic energy (i.e., velocity) resulting in a decrease in the gas transmission rate. Figure 1.4. Effect of Kinetic Energy vs. Surface Potential Energy on the molecules inside the pore 1.8. Evaporation of Liquids The evaporation rate of liquids at different temperatures can be calculated based on the pressure of the vapor at that temperature. If water is taken into consideration in a closed system, the rate the vapor which comes in contact with a metal is expressed as the following by Langmuir?s equation of mass flux (m) [27]: 27 RT MPm ?2= Where, P is the vapor pressure, M is the molecular weight of gas, R is the gas constant and T is the temperature. Kinetic theory of gases also suggests that at equilibrium, the evaporation rate is equal to the above equation. In a system, where equilibrium is not reached, the rate of evaporation or condensation is described by the Hertz-Knudsen equation as follows: p P s s RT M RT MPw ?? 22 ??= Where, w is the rate of evaporation in g/(cm2.s), Ps is the saturation vapor pressure corresponding to the liquid surface temperature and ?p is the Schrage correction for the mass movement of the vapor. Ts is the surface temperature of the liquid and Tp is the absolute temperature in the vapor region. If the difference in temperatures are ignored, and Ps- ?p =dP, this equation can be written as follows and used to calculate the rate of evaporation w in mass per unit area for a unit time: RT MdPw ?2= This assumption would be valid for most common liquids and can be used as an approximation [27]. Johnston et al. [26] studied the transport of water vapor through membranes and used the following equations for the rate of evaporation: )( )( airoutbl outinm Akw AkNAw ?? ?? ?= ?== 28 Where, N is the water flux, A is the membrane surface area, km the membrane mass transfer coefficient, ?in the density of water vapor in equilibrium as a function of water temperature, ?out the density of water vapor at the surface of the membrane, kbl the air boundary layer mass transfer coefficient and ?air the density of water vapor in the bulk air which flows through the membrane. 1.9. Microscale Models for Permeation of Gases through Protective Materials Gas flow and heat or mass transfer of protective clothing materials can be studied at three different scales. At a macroscale, protective clothing performance can be studied for the whole human body. At a mesoscale, a single limb which is covered by clothing can be evaluated and at a microscale the flow through individual fibers of the clothing material can be studied [6]. Microscale studies are important in understanding the transport behavior of different materials as well as the transmitted fluids and the effect of external conditions on transport. Although, these studies can be done experimentally or by modeling, modeling makes it easier to evaluate the transport behavior of different gases and surfaces and provide the basis for the selection and design of materials to be used in protective clothing. 1.9.1. Mathematical models for fibrous materials Many studies have been conducted to understand how porosity affects gas transmission through fabrics and membranes. In a study done on a perforated metal plate, it was shown that when pore diameters are constant throughout the thickness, (i.e. when tortuosity is 1), a large number of small pores allow higher water vapor permeation than a small number of large pores for the same porosity [58]. This study was also applied to 29 fabrics at a null pressure difference and the same relationship was validated for moisture vapor transport. However, for air permeation, experimental studies showed that permeation rate was inversely related to pore diameter and larger pores had a higher flow of air in fabrics [35]. The contradicting findings in those studies may be a result of the effect of moisture diffusion through the non-porous areas of the fabric in addition to the pores versus the convective flow of air through the openings which is usually dominant in comparison. It is also possible to deduce from these studies that gas permeation cannot simply be related to porosity alone and that other parameters such as pore diameters, pore distributions, pressure, temperature, tortousity, concentration, etc. also play an important role. McCord et al. [35] studied permeability through woven and knitted fabrics, which are significantly open structures compared to polymeric membranes. They considered the diffusion through the fibers negligible as the flow through the pores is very large in comparison. They created a model using Fick?s law of diffusion and Whelan?s model to calculate moisture and gas permeation through fabrics. Gibson et al. [19] studied transport properties of porous textiles of different fiber types using Darcy permeability and calculated flow rate as a function of pressure drop across the sample. They also derived an equation using the continuity equation and compared theoretical and experimental results. Water vapor transport was also studied in fabrics as well as microporous PTFE membranes using similar numerical models [21]. Another study on fibrous porous materials such as pulp and paper was done by Nilsson and Stenstrom [38] to understand gas diffusion through fibrous sheets. In this study, the relationship between gas diffusivity and volume fraction of cylindrical and band-shaped fibers was examined. 30 A general deduction from this study was that during water vapor diffusion, the main transport process occurred in the pores between the fibers rather than through the fibers by adsorption or absorption of gas molecules. Ryu et al. [42] also studied the gas transport through electrospun nonwoven mats made of Nylon 6 and the effect of pore sizes and surface area. It was shown that mass flow rate through the electrospun mats increased with increasing pressure for N2. However, increasing the concentration of the polymer increased the fiber diameter and therefore, increasing the polymer concentration on the fibers had an effect on permeability. For higher polymer concentrations, the gas flow rate was also lower compared to low polymer concentrations. This study showed that fiber diameters in electrospun non-woven mats had an effect on gas permeability as well as pore size and surface area and permeability mainly depended on fiber diameter and pore size. 1.9.2. Mathematical models for membranes Guijt et al. [22] used the Dusty Gas Model to predict the gas transport properties for a single component gas diffusion and compared with experimental results. In this study, all pores were assumed to be cylindrical and membrane morphology parameters were used for theoretical estimation of Knudsen, molecular and viscous flow. The results did not follow experimental data and it was necessary to evaluate membrane morphology parameters for molecular diffusion experimentally. Dusty Gas Model and Mean Transport Pore Model were also used by Hejtmanek et al. [23] to understand the effect of pressure build-up for multicomponent gas transport on porous media. The pressure change was largely affected by the molecular weight of the gases and the change was much larger for gas components with higher differences in molecular weights, i.e. 31 replacement of a heavier gas by a lighter gas would cause an increase in the pressure build-up inside the porous solid. This study found good correlation between experiments and the two models. Kerkhof et al. [31] developed a Velocity Profile Model versus the Dusty Gas Model due to the insufficiency of the latter model at small Knudsen numbers. This model was derived from Maxwell-Stefan equation with no-slip boundary condition for liquids and Maxwell-slip boundary condition for gases. The Velocity Profile Model presented a solution for transport of two-component liquids as well as gases through a single pore, for flow in one direction, ignoring radial transport. The chemical potential, and interspecies friction and shear stress is needed to predict the velocity profiles of any component. The results were found to be in good agreement with experiments for macro and mesopores. The results obtained from the Velocity Profile Model was also very similar to results with the Binary Friction Model which was presented by Kerkhof et al. [30] in a previous study for single component gases which included interspecies (diffusive) forces as well as species-wall forces. Gibson et al. [20] studied gas convection and water vapor diffusion properties in electrospun fiber mats. They suggested that when there is no pressure difference between the two sides of the mat, the water vapor transport occurs in the form of pure diffusion. In the case of the presence of a pressure difference, transport is carried out as a convective gas flow of the vapor. They used Darcy?s Law to calculate permeability of convective gas flow with the equation: ??? ? ??? ? ? ?? ? ?? ? ?= p x A Qk D ? 32 Where, kD is the permeability constant, ? is the gas viscosity, Q is the total volumetric flow rate, A is the sample area, ?x is the thickness, and ?p is the pressure drop across the sample. Shelekhin et al. [49] studied permeation, diffusion and separation of gases through molecular-sieve membranes. This study also described adsorption phenomena and created a mathematical model using the Dubinin-Radushkevich isotherm, which allows calculation of adsorption at different pressures and temperatures. The study showed that inorganic molecular-sieve membranes have better gas separation properties than polymeric membranes. Shindo et al. [50] used a different approach to study gas permeation through microporous glass membranes. In this study, the potential energy between the molecules and the pore walls was used to determine the permeation in the Knudsen regime. Since the collisions inside the micro pores in Knudsen regime occur only between gas molecules and pore walls and not between molecules, it is reasonable to study the permeation based on this relationship. Shindo et al.?s model was mentioned in section 7.1 in greater detail and this concept was also adapted in our research to understand the gas behavior inside the pores. Siginer and Bakhtiyarov [51] studied the flow of polymeric as well as Newtonian liquids in porous media mathematically and compared the pressure drop for both cases at different Reynolds numbers. They showed that at all Reynolds numbers, the pressure drop required for the same volume flow rate is higher for polymeric liquids than the Newtonian liquids such as water and water/glycerol solutions. Yoshioka et al. [59] studied gas permeation experimentally and theoretically for microporous silica 33 membranes with less than 1 nm pore size. Knudsen diffusion theory was used and flux calculations were carried out using Fick?s first law of diffusion. It was shown that as the pore diameter decreases, the effect of surface potential (i.e. interaction energy between the pore wall and the molecules) on permeation increases. In other words, as the pore size gets smaller, the surface potential in the pore begins to have a significant effect in permeation that cannot be ignored. 1.9.3. Molecular simulations and modeling Many researchers have studied flow based on behavior of individual molecules rather than using generalized gas laws. Petropoulos and Petrou [39] studied the permeation and adsorption of gases in pores and pore networks using a simulation model. They considered the surface of the adsorbent as a square lattice of adsorption sites on a 2- dimensional space. The z-axis, as the 3rd dimension showed the depth of the adsorption sites. Molecular trajectories were determined using the equations of motion. This study provided a much realistic theoretical estimation of the gas transport inside very narrow pores present in microporous membranes. Tunca and Ford [55] studied diffusion as well as adsorption through microporous membranes at a molecular level for steady-state permeation. They used free-energy calculations to define the adsorption sites on the adsorbent and used this information on a lattice model to simulate dynamic and equilibrium properties. They also used a continuum approach for diffusion calculations. Fick?s law of diffusion was applied to evaluate the diffusion flux and diffusion barriers between the adsorption sites. They used the model to simulate the permeation of methane through siliceous ZK4 and obtained reasonable estimations. This study was particularly useful to conclude that different 34 modeling techniques can be used to bridge the gap between molecular simulation and continuum transport modeling. Charrette and Houzelot [10] used both mathematical modeling and experimental techniques to study mass transfer through cotton fabrics. They studied the diffusion of salt through fabrics at zero pressure difference, using concentration as the driving force. They applied a mass balance to calculate the change of concentration in time. Then the fabric was washed in a washing machine to calculate the accumulation of NaCl as impurity concentration. This model was used to determine the efficiency of the washing process for a given fabric in certain conditions. 1.9.4. Adsorption studies Staudt et al. [53] studied the physisorption of pure gases such as He, O2, N2, Ar, CO2, etc. through microporous solids such as activated carbons or zeolites. Initially, single component gas adsorption measurements were conducted to evaluate the effect of gas composition, pressure and temperature on the amount adsorbed. Generally the absolute adsorbed amount can not be determined without the information regarding porosity, pore size distribution and the structure of the adsorbed layer. To eliminate this problem, Gibbs surface excess approach was adopted in this study which only requires the knowledge of void volume. The experiments were done gravimetrically to measure adsorption equilibria, by measuring the change of weight on the sorbent sample after adsorption of gas molecules. The experiments were found in good agreement with the three Parameter Isotherm (3-PIG) equation, which was developed to describe adsorption equilibria based on three parameters; maximum load, Henry?s law constant and the molar volume of the adsorbed phase. 35 Do et al. [14] studied adsorption in approximately 1 nm diameter pores of microporous solids and the effect of adsorbed molecules on the pore volume. It was suggested that in very small pores, the overlapping of adsorption potentials exerted by the opposite walls inside the pore may enhance the adsorption energy. Unlike larger pores which will only allow one layer of adsorbed molecules, smaller pores may have multiple layers due to this overlapping and the volume of adsorbed molecules may reach the volume of the pores. In this study, the following equation was suggested for the evaluation of the absolute amount adsorbed (nABS) at any given position r within the adsorption space: ?= v ABS dVrn )(? Where, ? is the density and V is the volume. The volume that should be used in this equation was also addressed by this study. Hu et al. [25] studied the effect of solubility of gases, free volume and sorption on gas permeability through dense membranes. Increasing the free volume caused an increase on permeability. It was assumed that gas sorption would have an effect on the permeation level. Since the increased free volume also caused an increase in solubility coefficient, to further investigate the effect of sorption, they produced membranes with the same free volume and different chemical structure. Studies done on polycarbonate (PC) and polyetherimide (PEI) membranes showed that even though PEI membranes had the highest gas sorption, they had the lowest permeability. PC membranes had the highest permeability, although they had the lowest gas sorption. Gas sorption did not seem to have an effect on the permeation order. 36 1.10. References [1] Adanur, S., ?Fabric Properties and Technology?, Wellington Sears Handbook of Industrial Fabrics, Technomic Publishing Co., Inc., Lancaster, PA, 1996. [2] ASTM D 1434-82: Standard test method for determining gas permeability characteristics of plastic film and sheeting, ASTM International, Reapproved 1998. [3] Attard, G. and Barnes, C., Surfaces, Oxford ; New York: Oxford University Press, 1998. [4] Barber, R. W. and Emerson, D. R., ?The influence of Knudsen number on the hydrodynamic development length within parallel plate micro-channels?, Advances in Fluid Mechanics IV, Eds. M. Rahman, R. Verhoeven, C.A. Brebbia, WIT Press, Southampton, UK, 2002, pp. 207-216. [5] Barker, R. L. and Scruggs, B.J., ?Evaluating the comfort performance of fabrics for nuclear protective apparel?, Performance of Protective Clothing: Fifth Volume, ASTM STP 1237, 1996. [6] Barry, J., Hill, R., Brasser, P., Sobera, M., Kleijn, C., Gibson, P., ?Computational fluid dynamics modeling of fabric systems for intelligent garment design?, MRS Bulletin, August 2003. [7] Bhandarkar, M., Shelekhin, A. B., Dixon, A. G., Ma, Y. H., ?Adsorption, permeation, and diffusion of gases in microporous membranes. I. Adsorption of gases on microporous glass membranes?, Journal of Membrane Science, Vol. 75, 1992, pp. 221-231. [8] Bird, R. B., Stewart, W. E., Lightfoot, E. N., Transport Phenomena, John Wiley & Sons, Inc., USA, 2002. [9] Carroll, T.R., Schwope, A.D., McCarthy, R.T., ?A technique to determine chemical contamination in chemical protective clothing?, ASTM Special Technical Publication, Fourth Volume, No. 1133, 1992. [10] Charrette, H., Houzelot, J. L., Ventenat, V., ?Modeling and experimental studies of mass transfer in cotton fabrics?, Textile Research Journal, Vol. 71, No. 11, 2001, pp. 954-959. [11] Chen, S. H., Yu, K. C., Houng, S. L., Lai, J. Y., ?Gas transport properties of HTPB based polyurethane/cosalen membrane?, Journal of Membrane Science, Vol. 173, 2000, pp. 99-106. [12] Cunningham, R. E. and Williams, R. J. J., Diffusion in Gases and Porous Media, 1980 Plenum Press, New York, 1980. 37 [13] Dhingra, S. S., ?Mixed gas transport study through polymeric membranes: a novel technique?, PhD Dissertation, June 19, 1997. [14] Do, D. D. and Do, H. D., ?Adsorption of supercritical fluids in non-porous and porous carbons: analysis of adsorbed phase volume and density?, Carbon, Vol. 41, 2003, pp. 1777-1791. [15] Faerevik, H., Markussen, D., Ogleand, G.E., Reinertsen, R.E., ?The thermoneutral zone when wearing aircrew protective clothing?, Journal of Thermal Biology, Vol. 26 , 2001. [16] Geankoplis, C. J., Transport Processes and Unit Operations, Englewood Cliffs, N.J.: PTR Prentice Hall, c1993. [17] George, S. C. and Thomas, S., ?Transport phenomena through polymeric systems?, Progress in Polymer Science, Vol. 26, 2001, pp. 985-1017. [18] Gibson, P. W., ?Effect of temperature on water vapor transport through polymer membrane laminates?, Polymer Testing, Vol. 19, 2000, pp. 673-691. [19] Gibson, P. W., Charmchi, M., ?Modeling convection/diffusion processes in porous textiles with inclusion of humidity-dependent air permeability?, Int. Comm. Heat Mass Transfer, Vol. 24, No. 5, 1997, pp. 709-724. [20] Gibson, P. W., Shreuder-Gibson, H. L., Rivin, D., ?Electrospun fiber mats: transport properties?, AIChE Journal, Vol. 45, No. 1, January 1999, pp. 190-195. [21] Gibson, P., Rivin, D., Kendrick, C., ?Convection/diffusion test method for porous textiles?, International Journal of Clothing Science and Technology, Vol. 12, No. 2, 2000, pp. 96-113. [22] Guijt, C. M. , Meindersma, G. W., Reith, T., de Haan, A. B., ?Method for experimental determination of the gas transport properties of highly porous fiber membranes: a first step before predictive modeling of a membrane distillation process?, Desalination, Vol. 147, 2002, pp. 127-132. [23] Hejtmanek, V., Capek, P., Solcova, O., Schneider, P., ?Dynamics of pressure build- up accompanying multicomponent gas transport in porous solids: intert gases?, Chemical Engineering Journal, Vol. 70, 1998, pp. 189-195. [24] Hoerner, S. F., ?Aerodynamic Properties of Screen Fabrics?, Textile Research Journal, Vol. 22, No. 4, 1952, pp. 274-280. [25] Hu, C. C., Chang, C. S., Ruaan, R.C., Lai, J. Y., ?Effect of free volume and sorption on membrane gas transport?, Journal of Membrane Science, Vol. 226, 2003, pp. 51- 61. 38 [26] Johnson, D. W., Yavuzturk, C., Pruis, J., ?Analysis of heat and mass transfer phenomena in hollow fiber membranes used for evaporative cooling?, Journal of Membrane Science, Vol. 227, 2003, pp. 159-171. [27] Jones, F. E., Evaporation of Water: with emphasis on applications and measurements, Chelsea, Mich. : Lewis Publishers, c1992, pp.26-38. [28] Kamada, K., Kamo, J., Motonaga, A., Iwasaki, T., Hosokawa, H., ?Gas permeation properties of conducting polymer/porous media composite membranes I?, Polymer Journal, Vol. 26, No. 2, 1994. [29] Keller, J. and Staudt, R., ?Gas Adsorption Equilibria: Experimental methods and adsorption isotherms?, Springer Science+Business Media, Inc, USA, 2005. [30] Kerkhof, P. J.A.M., ?A modified Maxwell-Stefan model for transport through inert membranes: the binary friction model?, The Chemical Engineering Journal, Vol. 64, 1996, pp. 319-343. [31] Kerkhof, P. J.A.M., Geboers, M. A. M., Ptasinski, K. J., ?On the isothermal binary mass transport in a single pore?, Chemical Engineering Journal, Vol. 83, 2001, pp. 107-121. [32] Krueger, G.P., ?Psychological and performance effects of chemical-biological protective clothing and equipment?, Military Medicine, Vol. 166 (12 Suppl), No. 41-3, Dec 2001. [33] Kumaran, V., ?Granular flow of rough particles in the high-Knudsen-number limit?, Journal of Fluid Mechanics, Vol. 561, 2006, pp. 43-72. [34] Luth, H., Surfaces and Interfaces of Solids, Springer-Verlag, Berlin, 1993. [35] McCord, M. G., Birla, K., Barker, R. L., ?The relationship between porosity and barrier effectiveness of some shell fabrics used in protective apparel?, Performance of Protective Clothing: Issues and Priorities for the 21st Century: Vol. 7, ASTM STP 1386, C. N. Nelson and N. W. Henry Eds., American Society for Testing and Materials, West Conshohocken, PA, 2000. [36] McGord, M. G., Brown, J. Z., Deaton, A. S., Barker, R. L., ?Comfort and barrier effectiveness of microporous films and laminates?, Performance of Protective Clothing: Seventh Volume, ASTM STP 1386, 2000. [37] McLellan, T. M., ?Work performance at 40 ?C with Canadian Forces biological and chemical protective clothing?, Aviation, Space and Environmental Medicine, Vol. 64, No. 12, December 1993, pp. 1094-100. [38] Nilsson, L. and Stenstrom, S., ?Gas diffusion through sheets of fibrous porous media?, Chemical Engineering Science, Vol. 50, No. 3, 1995, pp. 361-371. 39 [39] Petropolous, J. H. and Petrou, J. K., ?Simulation of molecular transport in pores and pore networks?, Journal of Chemical Society Faraday Trans, Vol. 87, No. 13, 1991, pp. 2017-2022. [40] Prutton, M., Introduction to Surface Physics, Clarendon Press, Oxford, 1994. [41] Rodriguez-Reinoso, F., Mc Enaney, B., Rouquerol, J., Unger, K., ?Characterization of porous solids VI?, Proceedings of the 6th Int. Symposium on the Characterization of Porous Solids, (COPS-VI), Alicante, May 2002, Studies in Surface Science and Catalysis, Vol. 144 Elsevier, Amsterdam etc., 2002. [42] Ryu, Y. J., Kim, H. Y., Lee, K. H., Park, H. C., Lee, D. R., ?Transport properties of electrospun nylon 6 nonwoven mats?, European Polymer Journal, Vol. 39, 2003, pp. 1883-1889. [43] Santee, W.R., Cadarette, B.S., Schamber, D.W., Gonzalez, R.R., ?Comparative responses to exercise-heat stress of two chemical protective garments?, Performance of Protective Clothing: Fourth Volume, ASTM STP 1133, 1992. [44] Satterfield, C. N., Sherwood, T. K., The Role of Diffusion in Catalysis, Addison- Wesley Publishing Company, Inc., Reading, MA, 1963. [45] Schofield, R. W., Fane, A. G., Fell, C. J. D., ?Gas and vapor transport through microporous membranes. I. Knudsen-Poiseuille transition?, Journal of Membrane Science, Vol. 53, 1990, pp. 159-171. [46] Schreuder-Gibson, H. L., Truong, Q., Walker, J. E., Owens, J. R., Wander, J. D., Jones Jr., W. E., ?Chemical and biological protection and detection in fabrics for protective clothing?, MRS Bulletin, August 2003, pp. 574-578. [47] Shalev, I., Barker, R. L., Hersh, S. P., Maini, S. M., Scruggs, B. J., Sood, V. K., Tosti, A., Tucker, P. A., Wu, G., ?Protective textile barrier systems and their comfort?, ASTM Special Technical Publication, No. 1237, 1996, pp. 347-359. [48] Shelekhin, A. B., Dixon, A. G., Ma, Y. H., ?Adsorption, permeation, and diffusion of gases in microporous membranes. III. Application of percolation theory to interpretation of porosity, tortuosity, and surface area in microporous glass membranes?, Journal of Membrane Science, Vol. 83, 1993, pp. 181-198. [49] Shelekhin, A.B., Dixon, A. G., Y. H. Ma, ?Theory of gas diffusion and permeation in inorganic molecular-sieve membranes?, AIChE Journal, Vol. 41, No. 1, January 1995, pp. 58-67. [50] Shindo, Y., Hakuta, T., Yoshitome, H., Inoue, H., ?Gas diffusion in microporous media in Knudsen?s regime?, Journal of Chemical Engineering of Japan, Vol. 16, No. 2, 1983, pp. 120-126. 40 [51] Siginer, D. A. and Bakhtiyarov, S. I., ?Flow in porous media of variable permeability and novel effects?, Journal of Applied Mechanics, Vol. 68, March 2001, pp. 312-319. [52] Sing et al., K. S. W., ?Reporting physisorption data for gas/solid systems with special reference to the determination of surface area and porosity?, (Recommendations 1984), Pure Applied Chemistry, Vol. 57, 1983, pp. 603-619. [53] Staudt, R., Herbst, A., Beutekamp, S., Harting, P., ?Adsorption of pure gases and mixtures on porous solids up to high pressures?, Adsorption, Vol. 11, 2005, pp. 379-384. [54] Stull, J. O., ?Considerations for design and selection of chemical-protective clothing?, Journal of Hazardous Materials, Vol. 14, 1987, pp. 165-189. [55] Tunca, C. and Ford, D. M., ?Coarse-grained nonequilibrium approach to the molecular modeling of permeation through microporous membranes?, Journal of Chemical Physics, Vol. 120, No. 22, June 2004, pp. 10763-10767. [56] Veghte, J.H., ?Field Test Evaluation of ASTM Standard F-1154 with Chemical Protective Suit Ensembles?, Performance of Protective Clothing: Fourth Volume, ASTM STP 1133, 1992. [57] Webb, P. A., ?Introduction to chemical adsorption analytical techniques and their applications to catalysis?, Micromeritics Instrument Corp. Technical Publications, January 2003. [58] Whelan, M. E., MacHattie, L. E., Goodings, A. C., Turl, L. H., ?The diffusion of water vapor through laminae with particular reference to textile fabrics?, Textile Research Journal, Vol. 25, No. 4, Mar. 1955, pp. 197-222. [59] Yoshioka, T., Nakanishi, E., Tsuru, T., Asaeda, M., ?Experimental studies of gas permeation through microporous silica membranes?, AIChE Journal, Vol. 47, No. 9, 2001, pp. 2052-2063. [60] Zhong, W., Ding, X., Tang, Z. L., ?Analysis of fluid flow through fibrous structures?, Textile Research Journal, Vol. 72, No. 9, 2002, pp. 751-755. 41 CHAPTER II GAS TRANSMISSION RATE THROUGH MICROPOROUS MEMBRANES Abstract Gas transmission through microporous membranes was investigated experimentally and a molecular-level probabilistic model was constructed to evaluate the effect of various parameters on the gas flow. Gas transmission experiments were conducted on PET, nylon and PP/PE microporous membranes at different pressure values and a comparison was made between permeability testing machines based on volumetric and manometric principles as to their ability to accommodate high permeability membranes. The interaction energy between gas molecules and membranes? surface was investigated to explore its influence on the gas transmission rate. Gas transmission experiments were simulated utilizing the proposed model, and results were compared to experimental data and to macro-level gas transmission models available in literature. Keywords: Air transmission, Microporous membrane, Probabilistic modeling 2.1. Introduction The idea of a garment that can protect first responders against chemical attacks or spills while providing comfort for a prolonged period of time is intriguing. Various garment material design constructions have been proposed in literature [13, 10, 12] with varied degrees of success. Selectively permeable layered systems [13] assembled from a group of microporous membrane layers and fabrics provide a logical premise to be able 42 to obstruct the passage of toxic gases while providing breathability to the skin. The first step to design such an assembly is to understand the process of gas permeation through single layer, and multi-layer assemblies, of microporous membranes. The permeation of gas molecules through a pore of a microporous membrane can be studied either at the level of individual molecules or at the level of the molecular population. At the molecular-level, each molecule and its interaction with other molecules and surfaces have to be taken into consideration and may be studied using a probabilistic technique. At the macro-level, the behavior of the molecular population depends on concentration and can be described by the continuum hypothesis. However, when the number of molecules in the molecular population decreases to a point where the behavior of the molecular population becomes chaotic, continuum equations cannot be used. The flux must be estimated by probabilistic techniques [2, 3]. The flow of any gas through a porous system is typically governed by geometric and energetic considerations. The geometric aspects are controlled by the mean free path of the gas (?), the average distance between collisions for a gas molecule, in comparison to the diameter of the pore [2, 3, 6, 4], and can be represented by the Knudsen number (Kn) as follows: DKn ?= Where, ? = mean free path and D = diameter of single pore. If the value of Kn is very small, the flow is considered to be in continuum (which can be considered as a viscous or a Poisseuille flow) assuming that the molecules move together with an overall average velocity and can be treated similar to a liquid flow inside 43 a tube [2]. In this case, Maxwell-Stefan and Navier Stokes equations can be used to adequately express the gas permeation process. As the value of the Knudsen number gets larger, the continuum assumptions breakdown [1], the gas no longer has an overall mass velocity or diffusion behavior and individual molecule-wall collisions become significant (Knudsen flow). Defining which regime should be followed to describe the gas transmission problem based on Knudsen number is not very clear. However, it is acceptable to consider the flow as either Knudsen if Kn>10 or viscous if Kn<0.01. The flow within the intermediate region of the Knudsen number interval is called the transition flow. Although most of gas flow in porous media lay in this region, there is a lack of analytical or numerical models to describe the transmission rate within this region. Moreover, these continuum regimes (Knudsen and viscous flow) do not consider the interaction between the gas molecules and the surface of porous media. Shindo et al. [11] considered the interaction energy between gas molecules and surfaces, during the flow of gases through pores, and compared it to the kinetic energy of the gas molecules. A gas molecule can either collide elastically with the pore surface or be trapped in the potential field and hop on the surface. If the kinetic energy of the molecule is larger than the interaction energy, that will make the molecule able to escape from the surface potential field and rebound elastically without losing energy, in other words, with the same velocity. Alternatively, if the interaction energy between the surface and the gas molecule is greater than the kinetic energy of the molecule, then the molecule will be trapped in the field, physisorbed on the surface and may cause a slip surface flow. The second case means that the molecule will lose some of its kinetic energy, and velocity, which results in a decrease of the gas transmission rate. 44 In this work, an attempt is made to develop a model to calculate the gas transmission rate based on simulating the random movement of the molecule inside the pore considering the molecule-surface physical interaction. To validate the model, gas transmission rate experiments using three different types of membranes were conducted and the results are compared to the model. 2.2. Experimental Work 2.2.1. Microporous membranes Three microporous membranes were used in these experiments: Poly(ethylene terephthalate) (PET) and nylon membrane manufactured by GE Osmonics Inc, and B130 membrane, a polypropylene/polyethylene (PP/PE) polymer blend, manufactured by BP. SEM images of these membranes are shown in Figure 2.1 and their physical properties are listed in Table 2.1. Table 2.1. Geometric properties of microporous membranes Membrane Name Porosity (%) Mean value of pore diameter (?m) Upper bound on pore diameter (?m) Lower bound on pore diameter (?m) Membrane thickness (?m) B130 0.046 1.28 2.87 0.6439 25 PET 9.40a 0.216 0.53 0.1846 10 Nylon 3.66 0.113 0.234 0.0839 90 a Provided by manufacturer 45 Figure 2.1. SEM images of microporous membranes: (a) PET (b) B130, and (c) nylon 46 Double layers PET membrane was also manufactured by covering one side of two single layer membrane by an epoxy compound and depositing unmodified TiO2 or SiO2 microparticles (~1 micron) on the coated surface and then the two layers were assembled under 39.0 x 105 Pa pressure at 140 oC for 30 min. Figure 2.2 shows SEM cross-sectional images for the assembled layers with incorporated microparticles which occupied about 20% of the volume between the two layers. Figure 2.2. Cross-section SEM images of double layer PET membranes with incorporated microparticles 47 2.2.2. Evaluation of gas transmission rate The transmission rate of air and ammonia through the three membranes, mentioned above, was measured at different pressure levels. Two machines were used to measure the air transmission rate, one based on the principle of manometric measurement (LYSSY L100-5000, manufactured by PBI-Dansensor), as shown in Figure 2.3, and the other on volumetric measurement (CSI-135, manufactured by Custom Scientific Instruments, Inc), as defined in the ASTM D 1434 standard, shown in Figure 2.4. Figure 2.3. LYSSY L100-5000, air permeability measurement chambers [7] Figure 2.4. CSI-135, view of the permeation cell [8] 48 The air transmission rate in the LYSSY L100-5000 manometric machine is measured based on the pressure difference between the upper and lower chambers and the time needed to increase the pressure from a set lower value to a set upper limit. The CSI-135 evaluates gas permeation by measuring the change of a volume of gas that permeates through the membrane from the lower to the upper chambers, at a constant pressure gradient. The change in volume is measured by the displacement of a short slug of mercury or other observable liquid in a glass capillary of a 1 mm diameter. Low porosity materials, or impermeable materials, require a high pressure difference value to test their permeability, while more open structures like microporous membranes require very low pressure. The manometric machine?s testing capabilities were limited to materials with very low permeability, therefore only the lowest permeability membrane, the B130, was tested with this machine. The standard pressure regulator on the CSI permeability instrument was more appropriate for less permeable or completely impermeable materials, therefore a modification was done to the machine to add a second pressure regulator for low pressure values. A similar modification to the pressure controls was not possible for the manometric machine. A digital camera was used to record the CSI experiments and the time needed for the slug to move a specific length was obtained by analyzing the product video. Five samples were tested at each pressure difference (?P) when evaluating the gas transmission rate (GTR) defined by the ASTM D1434 standard as ?the quantity of a given gas passing through a unit of the parallel surfaces of plastic film in a unit time under the conditions of test? and is calculated as follows in mol/(m?.s). GTR = TRA Vp ra **10 6? 49 Where, A = transmitting area of specimen (mm?), pa= ambient pressure (Pa), R = universal gas constant (R=8.314 L.Pa/(mol.K) and T = ambient temperature (K). Vr is calculated as Vr = slope*ac where ac = cross-sectional area of capillary (mm?) and slope = rate of rise of capillary slug (mm/s). 2.2.3. Evaluation of porosity and pore diameter A capillary flow porometer (CFP-1100-AEXS, manufactured by Porous Materials, Inc.) was used to evaluate the mean value and distribution of pore diameters (see Table 2.1). In this procedure, the pores of the sample are filled with a wetting liquid and air pressure is incrementally applied on one side of the sample. At a certain pressure, the larger pores are emptied of the wetting liquid and air flow starts. On further increase of the pressure, smaller pores are emptied, etc, and air pressure and flow through dry and wet samples is measured. Pore diameter is computed from the differential pressures between wet and dry samples. 2.3. Probabilistic Model A probabilistic model was developed utilizing a Monte Carlo approach to simulate gas transmission through microporous membranes in the CSI machine. The pore diameters were assumed to vary randomly between the upper and lower bounds observed experimentally and to be perfectly cylindrical. Pores were categorized based on pore size ranges and frequency and pore size averages were determined for each category. Molecules were assumed to fill available space, in the bottom chamber of the CSI machine, with random locations and direction of movement. The molecular trajectory was calculated for the molecules which were able to enter the pores. The molecules that 50 entered the pores travelled a distance ( nl ) assumed to be equal to the mean free path (?) or less in the case that the molecule hits the pore wall. The direction of the molecule (angle ?) after each collision was determined and the total path of the molecule inside the pore was calculated by following this trajectory. A sketch of a possible trajectory is shown in Figure 2.5. The total length of the path is equal to ? n i nl where n is reached as z = ? (thickness of the membrane). Figure 2.5. The sketch of a possible molecular trajectory The volume of molecules required to pass through the membrane at a certain time is calculated based on the volume of the capillary used in the experimental setup of the CSI machine. It is assumed that for each molecule that passes through the thickness of the membrane sample, a molecule will move into the capillary. The time (t1) needed for one molecule to pass through the pore is equal to: 51 ? pathmt = 1 Where, mpath = length of molecular trajectory path and v is the average molecule velocity, M RT ?? 8= . Matlab? software was used to construct the model. Model results were compared to experimental results, as well as gas flux calculated for Knudsen (JK) and viscous (JP) flow regimes mathematically defined as follows [9]: ????? /}8{32 2/1 PKPRTMMRTrJ KK ???= ????? /181 2 PKPRTMPrJ PP ???= Where, ? = porosity, r = average pore diameter, ? = tortuosity, ? = thickness of membrane, ? = viscosity, M = molecular weight, T = temperature, P = pressure and R = gas constant. 2.4. Results and Discussion Air transmission rate throygh B130 microporous membrane, which has the lowest porosity, was evaluated on both the manometric and volumetric permeability machines at different pressure-difference values. Air permeation rate reported by the manometric tester was calculated as 2.086*10-4 mol/m?s at a ?P=11.39 in-water, while reported as 3.32*10-5 mol/m?s on the volumetric tester CSI-135 at ?P=0.43 in-water. Linear extrapolating the data of CSI-135, assuming a linear relationship between GTR and ?P, showed that GTR will be 7.99*10-4 mol/m2s at ?P=11.39 in-water. Although results were 52 somewhat comparable, CSI-135 estimated a higher permeation rate for B130 than the manometric machine which may be a result of linear extrapolation based on very low pressure values. Nevertheless, it can be concluded from this comparison that the manometric machine has a limited pressure range and does not allow testing porous materials with high permeation rates, while CSI-135 is flexible in accommodating different types of materials due to its wide pressure range ability and smaller sample area. All further experiments for gas transmission rate were conducted on the CSI-135 machine. Figure 2.6. Air transmission rate of microporous membranes In air transmission experiments through untreated single-layer microporous membranes, shown in Figure 2.6, PET microporous membranes exhibited a high level of permeation followed by the nylon microporous membrane, while the B130 showed the lowest air transmission rate value. As expected, the increase in pressure caused an increase in the gas transmission rate for all three membranes. The error bars show the 53 spread of experimental results at a given pressure, possibly caused by the variability in the porosity of different membrane samples tested. The same behavior can be seen in Figure 2.7 for the transmission rate of ammonia through the three membranes. However, it was found that the transmission rate of the ammonia is higher than that of air because of its lighter molecular weight. Figure 2.8 shows, as an example, the transmission rate of both air and ammonia through a PET microporous membrane. Figure 2.7. Transmission rate of ammonia through microporous membranes 54 Figure 2.8. Comparison of air and ammonia transmission rates through PET A single layer of PET membrane covered with a thin layer of an epoxy compound was also tested for air permeability. The results showed a 14% reduction in the air transmission rate compared to the single layer untreated PET membranes, possibly due to blocking part of the pores? area. Two layer assembly of PET membranes were tested for air transmission rate and compared to single untreated PET membrane for each type of microparticles. Results are shown in Figure 2.9. As expected, single layer PET material exhibited a higher transmission rate, compared to both two-layer samples. The reduction in air transmission rate using double layer membrane was 50 to 70% less than the single layer at different pressures. A significant difference was not observed between membranes assembled with TiO2 and SiO2 spacers at lower pressures, however, as the pressure increased, a slight variation was visible. SiO2 had a comparatively lower permeation rate than TiO2 at higher pressures. As SiO2 was observed to cover more of the 55 open volume between the membranes when compared to TiO2, a slightly lower permeation rate than TiO2 was expected. Figure 2.9. Air transmission rate through double layered PET membranes Experimental data such as pressure values, pore diameter distribution, volume of permeation chamber, temperature, porosity, membrane thickness and sample size were used to model the experiment. The number of molecules within the confinement of the bottom chamber at the experimental pressure and temperature was calculated and the membrane surface was simulated based on the random pore distribution for a given porosity. The mean free path of the air molecule was calculated as the average value of the mean free paths of its constituents. Simulations were conducted for all three membranes and results were compared to Knudsen and viscous flow equations mentioned at the end of Section 2.3 and experimental data. Gaussian? molecular modeling software [5] was used to calculate the interaction energy between the air molecules and the surface of PET microporous membrane. Energy 56 calculations were restricted to the two most dominant components in dry air - nitrogen and oxygen as they represent almost 99% of the air volume (~78% and ~21% respectively). Two surfaces were modeled for the energy calculations - untreated PET microporous membrane and that covered with a thin layer of epoxy. Molecular models were geometrically optimized and the interaction energy calculated using the Density Functional Theory (DFT) using B3LYP-method with the base function (6-31g(d)). The interaction energy was calculated at different distances between the gas molecule and the surface to obtain an interaction energy profile. The depth of the energy-well was compared to the kinetic energy of the gas molecule calculated at 25oC to be 0.89 kcal/mol. Figure 2.10 shows the interaction energy profiles of both gases with virgin PET as compared to their kinetic energy. It can be seen from this graph that the molecules have enough kinetic energy to rebound from the surface. However, for surfaces covered with epoxy the interaction energy increased as shown in Figure 2.11. Although the nitrogen molecules still have greater kinetic energy to escape the surface, the oxygen energy-well is deeper than the kinetic energy. This means that if an oxygen molecule is 3.5 ? close to the surface it will be captured by the potential field of the surface (i.e., physisorped). It is important to note that although this represents a possibility of physisorption, the double layer specimens used in the experimental work shown in Figure 2. 11 were already in the air prior to the experiments allowing physisorption sites to be occupied by the oxygen molecules and limiting the contribution of physisorption to their GTR. 57 Figure 2.10. Interaction energy profiles for air and ammonia with PET Due to the large number of pores on the membrane surface involved in the simulation, a calculation-time barrier on the possibility of running a full-scale experiment was obvious and a scaling step was necessary. To ensure accuracy, a convergence study was conducted to evaluate the proper scaling parameters that will not affect the accuracy level of the analysis. Accordingly, the number of pores, created randomly within pore distribution obtained from the porosity test, was scaled to a smaller value. The calculations were done at this scale and then converted to the full scale based on the scaling ratio. 58 Figure 2.11. Interaction energy profiles for air and ammonia molecules and PET surface covered with a thin layer of epoxy Figures 2.12-2.14 show the comparison of the experimental data to the probabilistic model and the Knudsen and viscous flow equations for the PET, nylon and B130 microporous membranes, respectively. For PET, the average pore diameter (0.24?m) was much greater than the average molecular diameter of air (0.00037?m) and its average mean free path (0.0658?m). The Knudsen number was calculated to be 0.27 and the probabilistic results were within the Knudsen and viscous flow boundaries confirming the assumption of a transition regime for Kn >0.1. The probabilistic model showed acceptable correlation with the experimental values and was closer to a Knudsen flow. 59 Figure 2.12. Comparison of experimental data and numerical models for the permeation of air through PET membrane Figure 2.13. Comparison of experimental data and numerical models for the permeation of air through Nylon membrane 60 Figure 2.14. Comparison of experimental data and numerical models for the permeation of air through B130 membrane For the nylon microporous membrane, although the probabilistic model results were slightly higher than the experimental values, they were within acceptable limits. The correlation with experimental data was much better at lower pressures. As the pressure increased, the deviation increased slightly. Knudsen and viscous flow equations gave much lower estimations at the given pressures. Pore diameter of nylon is comparable to the average mean free path of air and Kn is calculated as 0.66. This large difference between the experiments and Knudsen and viscous flow equations might be a result of several approximations and assumptions which oversimplify the structure of microporous membranes with irregular shaped pores as seen in Figure 2.1. The nylon membrane 61 contained a large range of pore diameters and the calculated average pore diameter did not represent the real pore distribution. For the probabilistic results, pores were categorized based on pore diameter distribution and averages were calculated separately for each category during calculations. In Knudsen and Viscous flow equations, the average pore diameter was calculated for the overall structure, which caused further oversimplification of the real experimental conditions. Therefore, the probabilistic model enabled a more realistic estimation of the permeation rate, while Knudsen and Viscous flow equations result drastically deviated from the experimental values. In Figure 2.14, estimations based on probabilistic results were slightly higher than the experimental values and Knudsen and viscous flow equations estimated a higher permeation rate than the experimental results. While PET had a large pore population within a certain small range, B130 and nylon had a much wider range of pore diameters in large quantities. B130 is expected to display a comparatively higher viscous tendency based on Kn calculated as 0.05. It has the lowest Kn number compared to the other two membranes, therefore it is expected to be somewhat closer to a viscous type flow in comparison. However, with its low porosity value, B130 has a non-negligible amount of pores that are very close to the average mean free path of air, hence a Knudsen-like tendency is not completely unexpected. Figure 2.15 shows the experimental data for the transmission rate of ammonia compared to the probabilistic model results with and without the effect of the interaction energy between gas molecules and PET surface. An average of 9.4% reduction in permeation was observed as an effect of physisorption. 62 Figure 2.15. Comparison of experimental data for permeation of ammonia and the probabilistic model for uncoated PET microporous membrane, with and without the interaction energy between molecules and the surface The effect of membrane coating with an epoxy layer on gas permeation was studied utilizing molecular modeling for air (oxygen and nitrogen) and ammonia to model the interaction between the coat material and the permeate. As shown in Figure 2.11, nitrogen and ammonia molecules have enough kinetic energy to escape from the surface potential energy and elastically rebound. However, the oxygen energy-well is deeper than the kinetic energy which means that the molecule will be physisorbed on the surface if it comes 3.5 ? from the surface. The experimental data for the covered PET showed 14% reduction in the air transmission rate compared to a single layer untreated PET membrane. It is expected that since the membrane was subjected to air prior to the experiment that physisorption did not play a major role in this reduction. Instead, the cause may be a partial blockage of the area of the pore during coating. 63 The probabilistic model was tailored to calculate air transmission rate through double layer samples assuming that the pressure decreases linearly along the sample layers between the top and bottom chambers. Accordingly, the molecular separation distance and the number of molecules were calculated between the two layers. Probabilistic model results, calculated for particles occupying 20% of the volume between membrane layers, along with experimental data are shown in Figure 2.16. As shown in the figure, the probabilistic model results are very close to experimental data at smaller pressure values, however as the pressure increases, the deviation increases. Figure 2.16. Comparison of experimental results and model for double layered PET 2.5. Conclusions Gas transmission rate through three types of single layer microporous membranes and air transmission through one double layer membrane were evaluated. A probabilistic 64 computer-based model was developed to calculate the gas transmission rate through different types of microporous membranes and results were compared to experimental data as well as Knudsen and Viscous equations. The probabilistic model simulated the random variation of the pore size distribution, unlike Knudsen and Viscous equations that use average pore size, which resulted in more realistic results when compared to the experimental data especially for membranes with a wide variation in pore distribution. Acceptable correlations were observed between the probabilistic model and the experimental data. Gaussian?, molecular modeling software, was used to evaluate the interaction energy between the air molecules and both untreated PET surfaces and PET covered with a thin layer of an epoxy compound. The interaction energy profiles were compared to the kinetic energy of the molecule and results of untreated PET showed that the kinetic energy of air molecules is large enough to allow the molecule to escape the surface potential energy field. On the other hand, PET membranes coated with an epoxy compound can cause oxygen molecules to be physisorbed on the surface. The probabilistic model was also tailored to calculate air transmission rate through double layer microporous membranes and results were compared to experimental data showing an acceptable agreement. 2.6. References [1] Barber, R. W. and Emerson, D. R., ?The influence of Knudsen number on the hydrodynamic development length within parallel plate micro-channels?, Advances in Fluid Mechanics IV, Eds. M. Rahman, R. Verhoeven, C.A. Brebbia, WIT Press, Southampton, UK, 2002, pp. 207-216. 65 [2] Bird, R. B., Stewart, W. E., Lightfoot, E. N., Transport Phenomena, John Wiley & Sons, Inc., USA, 2002. [3] Cunningham, R. E. and Williams, R. J. J., Diffusion in Gases and Porous Media, 1980 Plenum Press, New York, 1980. [4] Fried, J. R., Polymer Science and Technology, Englewood Cliffs, NJ: Prentice Hall, 1995. [5] Gaussian 03, Revision D.01, M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, J. A. Montgomery, Jr., T. Vreven, K. N. Kudin, J. C. Burant, J. M. Millam, S. S. Iyengar, J. Tomasi, V. Barone, B. Mennucci, M. Cossi, G. Scalmani, N. Rega, G. A. Petersson, H. Nakatsuji, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, M. Klene, X. Li, J. E. Knox, H. P. Hratchian, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, P. Y. Ayala, K. Morokuma, G. A. Voth, P. Salvador, J. J. Dannenberg, V. G. Zakrzewski, S. Dapprich, A. D. Daniels, M. C. Strain, O. Farkas, D. K. Malick, A. D. Rabuck, K. Raghavachari, J. B. Foresman, J. V. Ortiz, Q. Cui, A. G. Baboul, S. Clifford, J. Cioslowski, B. B. Stefanov, G. Liu, A. Liashenko, P. Piskorz, I. Komaromi, R. L. Martin, D. J. Fox, T. Keith, M. A. Al-Laham, C. Y. Peng, A. Nanayakkara, M. Challacombe, P. M. W. Gill, B. Johnson, W. Chen, M. W. Wong, C. Gonzalez, and J. A. Pople, Gaussian, Inc., Wallingford CT, (2004) [6] George, S. C. and Thomas, S., ?Transport phenomena through polymeric systems?, Progress in Polymer Science, Vol. 26, 2001, pp. 985-1017. 66 [7] Operating manual of automatic manometric gas permeability tester, Model L100- 5000. [8] Operating manual of gas permeability cell system, Model CSI-135. [9] Schofield, R. W., Fane, A. G., Fell, C. J. D., ?Gas and vapor transport through microporous membranes. I. Knudsen-Poiseuille transition?, Journal of Membrane Science, Vol. 53, 1990, pp. 159-171. [10] Schreuder-Gibson, H. L., Truong, Q., Walker, J. E., Owens, J. R., Wander, J. D., Jones Jr., W. E., ?Chemical and biological protection and detection in fabrics for protective clothing?, MRS Bulletin, August 2003, pp. 574-578. [11] Shindo, Y., Hakuta, T., Yoshitome, H., Inoue, H., ?Gas diffusion in microporous media in Knudsen?s regime?, Journal of Chemical Engineering of Japan, Vol. 16, No. 2, 1983, pp. 120-126. [12] Stull, J. O., ?Considerations for design and selection of chemical-protective clothing?, Journal of Hazardous Materials, Vol. 14, 1987, pp. 165-189. [13] Truong, Q. T., Wilusz, E., Rivin, D.,? Advanced Lightweight Chemical/Biological Agent Protective Clothing Systems-Strengths and Limitations?, Membrane & Separation Technology News, September 2004. 67 CHAPTER III TRANSPORT BEHAVIOR OF MIMICS of WARFARE AGENTS THROUGH POROUS MEMBRANES Abstract Permeation of chemical warfare agents through microporous membranes were studied using their benign simulants. Gas transmission rate of these simulant gases (mimics) was investigated experimentally and an evaporation model was constructed to evaluate the evaporation rate of the mimic in liquid form and the effect of various parameters on the flow of vapor through the membrane. Beta-pinene and prenol were used as mimics of sarin and mustard gases, respectively. Experiments were conducted on PET and nylon microporous membranes at different temperatures. These transmission experiments were simulated utilizing the proposed evaporation model, and results were compared to experimental data. 3.1. Introduction Selectively permeable microporous membranes have been suggested as effective barrier materials against toxic gases when assembled with other membranes or fabrics [8, 9]. These assemblies can provide protection against harmful molecules by the obstruction created by the micropores, while they offer breathability to the skin for the comfort of the wearer. To design such assemblies, it is necessary to understand the transport process during permeation of toxic gases through such systems. In this study, it was assumed that 68 permeation of gas molecules occurs through the pores of the membrane, ignoring diffusion through the membrane material. Therefore, the level of permeation through such protective assemblies is mainly determined by the amount of porosity of membranes. The passage of molecules through the pores can be studied at two levels. (i) At a molecular level: When gas transmission process is examined at the level of individual molecules, the behavior of each molecule and its interaction with other molecules and surfaces is random and may be studied using probabilistic techniques. (ii) At a macro level: When the transmission process is examined at the level of molecular population, the behavior of the molecular system as a whole depends on the concentration of the permeate and can be described by the available gas laws [3, 4]. In this work, a macro level model is proposed. There are geometric and energetic considerations that govern the flow of any gas through a porous system. The geometric aspects are controlled by the mean free path of the gas (?), which is the average distance between collisions for a gas molecule, and the Knudsen number (Kn), a representative value for the mean free path in comparison to the diameter of the pore and is expressed as follows [3, 5]: DKn ?= Where, D is the pore diameter. These two values can be used to evaluate the nature of the gas flow and determine what kind of flow distribution should be expected. For very small values of Kn, the molecules move together with an overall average velocity and can be treated similar to a 69 liquid flow inside a tube [3]. In this case, Maxwell-Stefan and Navier Stokes equations are adequate to evaluate the permeation characteristics. As the value of the Knudsen number gets larger, the continuum assumptions break down, the gas no longer has an overall mass velocity or diffusion behavior and individual molecule-wall collisions become significant (Knudsen flow) [2]. Although Kn, by itself, cannot be used to determine the type of flow, it is acceptable to consider the flow as either Knudsen if Kn>10 or viscous if Kn<0.01 [2]. Most of the permeation models use generalized gas laws to estimate permeation rates, while some use molecular simulations and look at movements of individual molecules in the gas flow. This study combined both approaches and investigated the random behavior of molecules inside the gas flow as well as using generalized gas laws to estimate permeation through membranes, to obtain a better understanding of the transport process based on actual conditions of the flow. Knudsen number was used in this study to evaluate the nature of the flow and as a reference to determine and construct the appropriate permeation model to be used to estimate the transmission rate. 3.2. Materials Due to the highly toxic nature of warfare agents, they cannot be used in experimental settings in a laboratory without major precautions. Therefore, it is important to select benign chemicals as simulants of warfare agents, which are comparably similar to these toxic gases in molecular size, shape, composition or other physical properties. For this purpose, mimics of chemical warfare agents such as mustard and sarin have been obtained for this study in a liquid form and evaporated to study them as gases. 70 Mimics were mainly selected from chemicals used in flavor and perfume industry for food and cosmetics which were tested and approved regarding their benign character in case of interaction with human body. Information obtained from Flavor and Essence Manufacturers Association (FEMA) and Research Institute for Fragrance Materials (RIFM) regarding the toxicity of these chemicals confirmed that these chemicals had ideal properties for this type of studies and did not cause any harm to the researchers due to their tailored applications in foods or perfumes, their evaporation properties and average lifetime in their environment. Selected mimics included beta-pinene, a molecule with an almost spherical shape and can be used as a mimic of organophosphoric compounds such as the toxic sarin agent. Its ?ball-like? shape is also comparable to sarin in geometry. Its molecular weight and molecular diameter are also comparable to sarin as shown in Table 3.1. Other mimics selected for use in this study was methional and 3-methyl-2-butene-1-ol (prenol). Methional was found naturally in vegetables, such as tomatoes, and used as flavoring agent in food industry. Prenol also has a fruity odor and is used in the perfume industry. These mimics both have a ?stick-like? shape which is comparable to mustard gas and some of its derivates. The molecular diameters are also comparable to mustard as shown in Table 3.1. The vapor pressures of these mimics were also compared to the warfare agents during selection. Vapor pressure of ?-pinene and prenol are shown in Figure 3.5 and 3.6, respectively. The molecular morphology of toxic gases and mimics were drawn and molecular diameters were calculated using a Pymol (2005 DeLano Scientific LLC) software program as shown in Figure 3.1-3.4 and listed in Table 3.1. respectively. 71 Figure 3.1. Sarin Gas Figure 3.2. Beta-pinene (mimic for sarin) 72 Figure 3.3. Nitrogen Mustard Figure 3.4. Prenol (3-methyl 2-butene-1-ol) (mimic for mustard) 73 Table 3.1. Molecular diameter of air components, toxic agents and their mimics Molecule Molecular Formula Molecular diameter (A?) Molecular weight (g/mol) Description Sulfur Mustard C4H8Cl2S 8.00 159 Warfare agent Nitrogen Mustard C6H12Cl3N 8.00 204.5 Warfare agent Prenol C5H10O 7.30 86 Mimic gas for Mustard Methional C4H8OS 7.22 104 Mimic for Mustard Sarin C4H10FO2P 6.96 140 Warfare agent Beta-pinene C10H16 8.46 136 Mimic gas for Sarin Temperature-vapor pressure curves were obtained from literature for ?-pinene and prenol to guide in the evaluation of their pressure at different temperatures (Figure 3.5- 3.6) [7, 10]. 74 Figure 3.5. Vapor pressure of ?-pinene [7, 10] Figure 3.6. Vapor pressure of prenol [7, 10] Two microporous membranes were selected to carry out the permeation experiments: Poly(ethylene terephthalate) (PET) and nylon manufactured by GE Osmonics Inc. These 75 membranes had different physical properties as listed in Table 3.2. PET was hydrophilic, while nylon was hydrophobic and they had different surface structures as shown by the SEM images in Figure 3.7. These differences in material properties allowed comparison of the permeation behavior on these membranes. Table 3.2. Geometric properties of microporous membranes Membrane Name Porosity (%) Mean value of pore diameter (?m) Upper bound on pore diameter (?m) Lower bound on pore diameter (?m) Membrane thickness (?m) PET 9.40a 0.216 0.53 0.1846 10 Nylon 3.66 0.113 0.234 0.0839 90 a Provided by manufacturer (a) (b) Figure 3.7. SEM images of microporous membranes: (a) PET (b) nylon 3.3. Experimental Work The volumetric permeability instrument CSI-135 (manufactured by Custom Scientific Instruments, Inc) based on ASTM D1434 standard was used to measure permeation of 76 mimics through microporous membranes. It evaluates gas permeation by measuring the change of a volume of gas that permeates through the membrane from the lower to the upper chambers, at a constant pressure gradient. The change in volume is measured by the displacement of a short slug of mercury or other observable liquid in a 1 mm diameter glass capillary. Mimics were obtained in the form of a volatile solution at room temperature. The CSI machine was modified to allow testing of an evaporating liquid by closing the gas feeding pipes in the lower chamber and allowing control of the temperature inside it. Insulated thermocouple wires with 0.4 mm thickness and a digital temperature indicator were used to monitor the temperature of the mimic during the experiments in the bottom chamber as shown in Figure 3.8. Thermocouple wires were passed through the vent line of the cell bottom chamber to be in continuous contact with the mimic during the test. Results showed that temperature of the mimic was constant during the experiment with a fluctuation in the range of ?0.5 ?C. Figure 3.8. Schematic of bottom chamber (left) and its image (right) with thermocouples wires 77 To achieve a better control on the temperature in the cell and the mimic inside it, several trials were conducted utilizing tape heaters and pre-heated oil bath. Both approaches gave high fluctuations in temperature due to air drafts and difficulty of temperature control in the immediate environment. To avoid these effects, the permeation cell of the CSI machine was placed in a controlled humidity and temperature cabinet (Figure 3.9), and the readings of temperature were taken from three different sensors placed outside as well as the thermocouple inside the cell. Experiments were conducted after these sensors matched, indicating thermal equilibrium between the mimic inside and the whole testing system. Figure 3.9. Modified setup for measurement of permeation of mimics To calibrate the CSI instrument for the vapor pressure of the mimic, the mimic was placed inside the bottom chamber, without the membrane between the chambers. The slug inside the capillary applies a pressure on the gas below, of which value is 78 proportional to its length. The slug remains steady at a point where the vapor pressure of the gas below is equal to slug pressure. The slug length (pressure) was calculated for that point and the vapor pressure was calculated based on literature values for the applied temperature. The measured vapor pressures were compared to literature values and a good agreement was achieved. Gas transmission rate (GTR) was calculated as follows in mol/(m?.s) [1]: GTR= TRA Vp ra **10 6? where, A = transmitting area of specimen (mm?), ap = ambient pressure (Pa), R = universal gas constant (R=8.314 L.Pa/(mol.K) and T = ambient temperature (K). Vr is calculated as Vr = slope*ac, where ac = cross-sectional area of capillary (mm?) and slope = rate of rise of capillary slug (mm/s). 3.4. Evaporation Model As previously mentioned, mimic gases are not available in large amounts in the form of a gas which can be fed through the permeation tester like other gases. Since they are obtained in the form of a volatile solution at room temperature, their temperature must be increased to obtain vapor. To model transmission of mimics through membranes, it is necessary, as a first step, to understand the evaporation behavior of the mimics at high temperatures. The important parameters to consider are: temperature of the mimic, pressure of the evaporating mimic inside the bottom chamber, the evaporation rate, the changing 79 diameters of the vapor passageway structure in the permeation setup and the relationship between the molecule diameter of the mimic and the pore diameter of the membrane. The molecule diameters of mimics and the average pore diameters of microporous membranes were used to calculate and determine the flow regime that should be expected. Knudsen number for a mimic such as ?-pinene was calculated to be around 0.09. Since the permeation of the mimic in liquid form depends mostly on its vapor pressure and evaporation rate and the Knudsen number falls closer to a slip flow category, a single molecular velocity could not be used for modeling. A macro-level behavior, i.e. the behavior of the molecular population, was taken into consideration rather than single molecular movement inside the pores. Evaporation rate (mass loss rate) was calculated in kg/s by adopting the evaporation rate equation for liquids: RT MPambPv UnitArea teMassLossRa ?2)( ?= (1) Where, Pv = vapor pressure, Pamb = ambient partial pressure and M=molecular weight [6, 11]. Since the vapor pressure inside the mimic never reaches its saturated value, the difference between the vapor pressure and ambient partial pressure of the mimic (dP) was determined through experimental observations. As vapor pressure increases with temperature, this difference was assumed to increase and the ratio of increase was correlated to the ratio of increase in vapor pressure based on values found in literature for each mimic. 80 It was also observed in the existence of a membrane between the two chambers that evaporation starts before the mimic would reach its full pressure value under the membrane. Experiments were conducted to understand the behavior of evaporating gases to support the modeling efforts. Mass loss rate for evaporation was measured for ?-pinene with and without the PET membrane within a temperature and humidity controlled environment. Results showed a dependence of the evaporation on pressure build-up under the membrane prompting further experimental verification of the true conditions inside the testing device. In a typical permeation experiment, where the pressure on either side of the slug is equal and the pressure difference is zero, the slug inside the capillary sits at a resting position as shown in Figure 3.10a. As the slug starts to move up from that initial symmetric position, the pressure difference begins to build up based on slug height as shown in Figure 3.10b. The pressure of the slug can be calculated as Ps=?gh, where ?=density of the slug material, g=gravity and h=slug height. When the slug reaches its full length beyond its initial position, the pressure difference reaches its full value (Figure 3.10c) and is equal to Ps. By monitoring the movement of the slug versus time using a digital video camera and calculating the pressure based on length of the slug, it was observed that the pressure required to move the slug was much lower than the vapor pressure that was expected to be reached at the given temperature. The reason is that; as the vapor accumulates and starts increasing the pressure P inside the chamber, any value of P that is slightly above Ps, i.e. any value of P where P-Ps>0, will be sufficient to move the slug. The movement of the slug was observed to be at a constant rate as shown in 81 Figure 3.11. Hence the rate of evaporation was calculated based on the pressure required to move the slug rather than the expected vapor pressure. Figure 3.10. Slug movement versus Pressure relationship y = 1.30E-02x + 6.10E-01 0 10 20 30 40 50 60 70 80 90 1200 2200 3200 4200 5200 6200 7200 Time (s) dP (P a) Figure 3.11. Pressure versus time during slug movement due to evaporation of mimics Another experiment was conducted to support the modeling effort, to evaluate the effect of the existence of an orifice on the evaporation rate for a given volume of ?- pinene inside a closed and open beaker at the same temperature levels. Since the gas is traveling through a sample of 1 inch diameter and reaching a capillary of 1 mm diameter, (a) (b) (c) 82 the diameter of the opening is decreasing by a ratio of 0.1/2.54=0.039 cm as shown in Figure 3.12a. Figure 3.12. Orifice structure comparison between permeation cell and beakers Six beakers of 4 cm diameter were filled with 2 grams of ?-pinene. Three of them were kept open while the other three were covered with an aluminum foil. A small hole of 0.157 cm diameter was punched on the aluminum foils (Figure 3.12b), which was comparable in area to the ratio of the opening on the top chamber in CSI-135 permeability tester, to examine the rate of evaporation in terms of mass loss-rate of ?- pinene. It was observed that the mass loss-rate in open beakers was drastically higher than that for beakers with a hole and with a ratio that is close to the ratio of the open area to the closed area. Therefore the drop caused by the decreasing orifice diameter was included in the evaporation rate calculations to show the decrease in mass loss rate. These experiments provided a better understanding regarding the determination of parameters such as the pressure difference in the system and the evaporation behavior of the mimic. Based on these understandings, the permeation model was refined to evaluate mimic permeation rates through membranes. (a) (b) 83 3.5. Results and Discussion 3.5.1. Experimental results Permeation experiments using the CSI-135 instrument were conducted for PET and nylon membranes with beta-pinene and prenol as a mimic. Figure 3.13. GTR vs. Temperature for ?-pinene The transmission rate of beta-pinene through PET and nylon is shown in Figure 3.13. For both membranes, increasing the temperature causes an increase in the permeation rate, since the vapor pressure is higher at higher temperatures, which causes a faster build-up of pressure to move the slug. Since PET has a higher porosity relative to nylon membranes, a higher transmission rate was expected for PET membranes. As the temperature increases, PET displays a linear increase in permeation rate while nylon 84 shows a nonlinear response. This may be due to the complex mesh-like structure of nylon which makes it possible for the membrane to absorb evaporating molecules at a faster rate in comparison to PET, which has straight cylindrical pores, where the permeation occurs at the regular rate of evaporation. Figure 3.14. GTR vs. Temperature for Prenol GTR of prenol through PET and nylon membranes is shown in Figure 3.14. At low temperatures, the transmission rate did not change much. As mentioned in the previous sections, prenol has a lower vapor pressure compared to beta-pinene at the same temperature (Figure 3.5-3.6). During the initial experiments with no membrane, it was also observed that the evaporation rate was significantly lower relative to beta-pinene. Therefore, it is not beyond expectations that the transmission rate does not change much with temperature as seen in experimental results. This is particularly true for the PET 85 membrane as shown in Figure 3.15. For the highest temperature points, a slight decrease in the transmission rate is observed; however, it is not possible to come to a conclusion that this is actually a drop since it may be a result of variability caused by the sensitivity of the experiments. There are several other parameters that affect the rate of transmission such as the molecular size and weight of the evaporating gas and the interaction of the mimic with the membrane which cannot be deducted from this graph alone. However, the effect of porosity can be seen on Figure 3.14, since PET shows a relatively higher transmission rate than nylon which is expected due to higher porosity of PET. Figure 3.15. Comparison of beta-pinene and prenol through PET When the permeation rate of the two mimics through PET membranes is compared, as seen in Figure 3.15, there is not an observable difference with the results at lower pressures. Prenol has a smaller molecular size and a lower molecular weight than beta- pinene and therefore it is expected to have a higher permeation rate. However, it has a 86 much lower evaporation rate and the vapor pressure is about 5 times higher than beta- pinene at the same temperatures. Therefore, the increase in permeation rate is low for prenol at small pressure differences. Beta-pinene, on the other hand, displays an increase in permeation rate with increasing temperature. Compared to other gases such as air and ammonia (Figure 2. 8), the permeation rate of mimics was significantly lower for both beta-pinene and prenol for all membranes. Since beta-pinene and prenol have a much larger molecular size compared to air or ammonia, it is expected to have a lower permeation rate for these mimics. In addition, beta-pinene and prenol come in liquid form and need much higher temperatures to evaporate which is a limit on the permeation process in comparison to ammonia. Even at the highest pressure values in which they could be tested, mimics displayed a significantly lower permeation rate. This shows that it is possible to develop protective materials that allow permeation of air, while permeation of toxic gases with high molecular diameter is significantly lower. 3.5.2. Comparison of experimental data to modeling results Based on experiments mentioned above, the pressure of the gas inside the bottom chamber was assumed to always increase up to a pressure that is enough to move the slug and the partial pressure of the gas would not reach the saturated vapor pressure value. By dividing the mass loss rate by the density of the gas, we can obtain the flow rate of the vapor inside the bottom chamber in m3/s. The flow rate will drop as the gas passes through the porous media and will drop further as it goes through the capillary due the difference in the cross-section area between the top chamber and the capillary as an orifice. Results based on this model and using equation (1) were compared to 87 experimental data for the permeation of ?-pinene and prenol through PET and nylon membranes as shown in Figure 3.16-3.19. Figure 3.16. GTR versus temperature for evaporation of beta-pinene through PET membrane Figure 3.17. GTR versus temperature for evaporation of beta-pinene through nylon membrane 88 Figure 3.18. GTR versus temperature for evaporation of prenol through PET membrane It can be seen from Figure 3.16 that the estimated permeation rate using the evaporation model has an acceptable correlation with permeation of beta-pinene through the PET membranes. The results are particularly close at low temperatures, while there is a higher deviation at higher temperatures. The linear increase of transmission rate obtained during the experiments was similar to the results obtained by the evaporation model which was based on the increase in temperature. On the other hand, for nylon as shown in Figure 3.17, the model estimated a lower permeation rate compared to PET, which was originally expected due to lower porosity and higher thickness of nylon relative to PET. However, the experimental results were higher and did not show a linear increase of the transmission rate with temperature unlike the model estimates. This could be a result of the high tortousity and mesh-like structure of nylon, which could not be considered in the model. 89 Figure 3.19. GTR versus temperature for evaporation of prenol with Nylon The model estimations for PET, shown in Figures 3.16 and 3.19, were comparable to experimental results. However, in the permeation of prenol through PET (Figure 3.18), the model showed a slight linear increase with the increase in temperature, unlike the experimental data. In the case of nylon, the transmission rate estimations were much lower than the experimental results for both mimics. For the permeation of beta-pinene (Figure 3.17), modeling results were very close to experiments at low temperatures; however, the experiments did not show a linear increase with increasing temperatures, contrary to model estimations. This can be a result of not accounting for the tortuous structure and the absorption capacity of nylon in the model. Model estimations for the permeation of prenol through the nylon membranes are shown in Figure 3.19. Prenol is expected to have a higher transmission rate due to smaller molecular weight and molecular size of the mimic, however, its low vapor pressure and evaporation rate has a more dominant effect and causes a lower transmission rate. The high deviation in Figure 90 3.19 can be a result of a combination of these mutually-exclusive effects of mimic properties and the complex morphology of the nylon membranes. Experimental results for mimics showed some variability at the same temperature, due to the complex nature of the vapor and the sensitivity of the experimental setup. However, it is expected that as the temperature increases the transmission rate of the gas will increase due to an increasing vapor pressure and higher evaporation rate. The evaporation model showed reasonable correlation with the experimental data with an expected increase of GTR at increasing temperatures. 3.6. Conclusions Gas transmission rate of mimics of warfare agents such as mustard and sarin was studied experimentally and analytically. A volumetric air permeability tester was modified to enable the measurement with mimics in liquid form. An evaporation model was developed to calculate the permeation rates of mimic vapors through microporous membranes at different temperatures based on the effect of evaporation rate of the mimic. Results were compared to experimental data and showed an increase in gas transmission rate with the increase in temperature. Acceptable correlations were observed between experiments and the model. It can be concluded from this study that permeability characteristics of membranes can be predicted more accurately for membranes with a well-known morphology. It is also possible to improve barrier properties of membranes for toxic gases with larger molecule diameters than air, while a higher level of air permeation is still available. 91 3.7. References [1] ASTM D 1434-82: Standard test method for determining gas permeability characteristics of plastic film and sheeting, ASTM International, Reapproved 1998. [2] Barber, R. W. and Emerson, D. R., ?The influence of Knudsen number on the hydrodynamic development length within parallel plate micro-channels?, Advances in Fluid Mechanics IV, Eds. M. Rahman, R. Verhoeven, C.A. Brebbia, WIT Press, Southampton, UK, 2002, pp. 207-216. [3] Bird, R. B., Stewart, W. E., Lightfoot, E. N., Transport Phenomena, John Wiley & Sons, Inc., New York, 2002. [4] Cunningham, R. E. and Williams, R. J. J., Diffusion in Gases and Porous Media, 1980 Plenum Press, New York, 1980. [5] Fried, J. R., Polymer Science and Technology, Englewood Cliffs, NJ: Prentice Hall, 1995. [6] Jones, F. E., Evaporation of Water: with Emphasis on Applications and Measurements, Chelsea, Mich. Lewis Publishers, c1992, pp.26-38 [7] Jordan, T. E., Vapor Pressure of Organic Compounds, New York, Interscience Publishers, 1954. [8] Stull, J. O., ?Considerations for design and selection of chemical-protective clothing?, Journal of Hazardous Materials, Vol. 14, 1987, pp. 165-189 [9] Truong, Q. T., Wilusz, E., Rivin, D., ?Advanced lightweight chemical/biological agent protective clothing systems: strengths and limitations?, Membrane & Separation Technology News, September (2004) [10] Yaws, C. L. Handbook of Vapor Pressure Vol. 3, Gulf Publications, 1994. 92 [11] Zemansky, M. W., Heat and Thermodynamics: an Intermediate Textbook for Students of Physics, Chemistry, and Engineering, New York, McGraw-Hill, 1951. 93 CHAPTER IV CHEMICAL ADSORPTION OF MICROPOROUS MEMBRANES COATED WITH ACTIVE ELEMENTS Abstract Chemical warfare agents and toxic industrial chemicals were reviewed and several toxins among these groups were selected for this study. Simulants of those toxins or benign equivalents were obtained to use in experimental evaluations of the transmission behavior of those chemicals through microporous membrane materials proposed for chemical protection. To enable blockage of the toxic molecules by the protective material surfaces, microporous membranes were treated with active elements which can interact with the toxic gas molecules and prevent them from proceeding through the material. The effect of physical and chemical adsorption of gases, such as ammonia, on gas transmission rate was evaluated experimentally for membranes made of PET coated with PGMA and activated with functional groups XL-I-177 and 5-azido-3- oxapentanesulfonate by anchoring them on PET surface. XL-I-177 was selected for its ability to interact with mustard gas or its mimic methional and 5-azido-3- oxapentanesulfonate for its ability to capture NH3. A probabilistic chemisorption model was developed to understand the transmission behavior of membranes in the presence of active elements. Interaction potential of the active surfaces with selected gas molecules were calculated using molecular modeling software Gaussian?. An experimental setup was developed to measure gas concentration changes upon passage through active 94 membranes and results were compared to concentration behavior of membranes with no active agents. Experimental results and probabilistic estimations were also compared to understand the adsorption phenomena. 4.1. Introduction Chemical Warfare Agents (CWA) and Toxic Industrial Chemicals (TIC) are toxins that can become a threat to human population in the case of exposure and cause death or irreparable damage to human body. CWA has been used often in recent history against civilians as well as military personnel. Due to their fast and lethal results, all of these toxins must be decontaminated or prevented from contact with the skin [2, 3]. CWAs can be categorized such as nerve gases (sarin, soman, tabun, etc.), blister agents (mustard gas, lewisite, etc.), chocking agents (phosgene, chlorine), vomiting agents (adamsite, etc.), blood agents (cyanide, cyanogen, etc.), tear gases and incapacitating agents. Each category of gases causes damage or failure to different parts of the body. TICs are usually categorized based on their toxic effect on human body and their physical characteristics such as flammability, explosiveness, etc. [9]. Gases investigated in this study were mustard, a blister agent; sarin, a nerve agent and ammonia a TIC with basic character which reacts violently with strong oxidants and halogens. The most common way to evaluate the performance of protective materials is to measure their permeation rate for different gases. This information is useful in selection of more resistant materials for protection. 95 4.2. Adsorption Phenomena Adsorption is defined as the attachment of gases or liquids onto a solid surface [4]. Unlike absorption, adsorption occurs on the surface and does not cause the gases or liquids to dissolve into another liquid or solid. Adsorption of molecules onto a solid surface depends strictly on their energy state close to that surface. When there is no energy loss, the molecules are considered to have an elastic collision with the surface without bonding with it. If there is a small energy loss, the collision is considered to be inelastic. When a sufficient energy loss occurs, adsorption takes place. Adsorption may occur in the form of physical adsorption (physisorption) or chemical adsorption (chemisorption). Physisorption occurs as a result of a physical interaction between the gas and the solid and does not cause a change in the electronic structure of the molecule. The gas and the solid are bonded with van der Waals forces which can occur on multiple layers. Chemisorption is usually a result of strong covalent or ionic bonds which requires higher adsorption energies and is not easily reversible. It usually happens on adsorptive sites as a monolayer [1, 6, 10]. The adsorption energy or the energy loss required for adsorption is dependent on the translational kinetic energy of the molecule (Ek) as well as the attraction between the molecule and the surface which is related to the surface potential energy (Ep). If the attraction between the molecule and the surface is higher than the kinetic energy of the molecule (Ek?Ep), the molecule is more likely to experience energy loss, i.e. lose its kinetic energy and become adsorbed by the surface. Energy loss of the molecules is a function of the distance of the molecule from the solid surface [7]. As the molecules get closer to the surface, the attraction energy 96 increases and causes the molecule to lose its kinetic energy. At the distance, where the kinetic energy becomes lower than the interaction energy between the molecule and the surface, adsorption takes place [8]. Since chemisorption requires higher adsorption energy, chemisorption of the molecule occurs at a closer distance to the surface than the distance required for physisorption [1, 10]. Therefore, when there is strong attraction between the gas and the solid surface, chemical adsorption has priority over physical adsorption [6]. In the evaluation of adsorption behavior of a system reaching equilibrium, Langmuir adsorption isotherm is most commonly used. For chemisorption, Langmuir isotherm equation assumes that there are only a fixed number of sites available for adsorption each of which can only be occupied by one molecule. At a constant temperature, there is a dynamic equilibrium between the adsorbed layer and the gas at pressure P. The adsorbate molecules are constantly colliding with the surface. If they collide with an available adsorption site, they are adsorbed. If they collide with a filled adsorption site, they bounce back into the gas phase without sticking on the surface. The fraction of coverage on the surface can be written as follows: N Nads=? Where, Nads=number of adsorbed molecules and N=available active sites. Where all sites are filled ?=1. At zero pressure, adsorption may be too slow and equilibrium may not occur easily, while at higher pressures, the equilibrium can be reached at a faster rate [1, 10]. 97 4.3. Membrane and Gas Properties A polymeric membrane made of PET with a pore size of 0.2 ?m, porosity of 9.4% and thickness of 10 microns was used in chemisorptions studies. The PET membrane was selected for the attachment of functional groups which can act as active elements and interact with the gas molecules to prevent them from proceeding into the membrane layers. Modifications of PET membrane was carried out at Clemson University in which plasma treated PET membranes were covered with an epoxy containing polymer poly (glycidyl methacrylate) (PGMA) to modify its surface and create an anchoring layer. To ensure that the pores on the membrane were not blocked by this treatment, PET grafted with PGMA was tested for air permeability. No significant blockage of the pores was observed during the experiments. PGMA modified layers were, then, used for attachment of the functional groups on the surface. A model chemical compound XL-I-177 (Figure 4.1) and another active element with an acidic functionality 5-azido-3-oxapentanesulfonate (Figure 4.2) were attached separately on the surfaces of PET microporous films. Figure 4.1. Chemical structure of XL-I-177 98 Figure 4.2. Adsorption of NH3 on modified PET with 5-azido-3-oxapentanesulfonate 4.4. Modeling of Chemical and Physical Adsorption To develop a permeation model which includes adsorption, it is necessary to understand the attraction mechanism between the gas and the solid and define the important parameters that affect adsorption and as a result affect the overall rate of permeation on materials created with active elements. The important parameters to take into consideration are the gas type, energy parameters between the surface and the molecule, pore structure of the membrane, density of the active elements on the surface and the number of molecules satisfying the adsorption conditions. At the initial step of constructing the model, the energy requirements were investigated and the minimum distances necessary for chemical and physical adsorption were calculated for the condition Ek?Ep. A geometric probabilistic approach was adopted to calculate the amount of molecules attached to the surface until an adsorption equilibrium state is reached. 4.4.1. Energy calculations Molecular geometry optimization and energy calculations were done using Gaussian? molecular modeling software for the toxic industrial chemical ammonia, warfare agent mustard and its mimic prenol to enhance the understanding of their adsorption mechanisms on the treated PET surface before and after being grafted with 99 functional groups. Calculations were done using the Density Functional Theory (DFT) using B3LYP-method with the base function (6-31g(d)). For PET treated with PGMA, it was shown that all three gases had a higher kinetic energy than the energy required for adsorption. Therefore a chemical attachment of these gases to the PET surface was not expected after PGMA treatment as shown in Figure 4.3. On the other hand, after the attachment of 5-azido-3-oxapentanesulfonate or XL-I-177 on the PGMA treated PET, the kinetic energy of all gases were found to be less than the required adsorption energy and therefore an attachment of the gases on the active surfaces was expected. The distance from the surface that is required for an adsorption to take place was calculated by measuring the intersection of the energy curve with the kinetic energy line for each gas molecule. For instance, in Figure 4.4, adsorption of ammonia by the active PET surface is expected when the molecule is at 4.9 ? or closer distance from the surface, but more than 2-3 ?. Figure 4.3. Interaction energy between the gases and the surface covered with PGMA -1 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 1 2 3 4 5 6 7 8 9 10Distance between the molecule and the surface (?) In ter ac tio n e ne rg y b etw ee n t he m ole cu le an d t he su rfa ce (k Ca l/m ol) NH3 Mustard gas 3-methyl-2-buten-1-ol Kinetic Energy of gas molecule 100 -12 -10 -8 -6 -4 -2 0 2 2 3 4 5 6 7 8 9 10Distance between the molecule and the surface (?) In ter ac tio n e ne rg y b etw ee n t he m ole cu le an d t he su rfa ce (k Ca l/m ol) Kinetic Energy of gas molecule Figure 4.4. Interaction energy between ammonia and the surface with active element (5- azido-3-oxapentanesulfonate) -4 -3.5 -3 -2.5 -2 -1.5 -1 -0.5 0 0.5 1 2 3 4 5 6 7 8 9 10Distance between the molecule and the surface (?) Int er act ion en er gy be tw ee n t he m ole cu le an d t he su rfa ce (k Ca l/m ol) Mustard gas 3-methyl-2-butenol (mimic gas) Kinetic Energy of gas molecule Figure 4.5. Interaction energy between mustard gas and prenol and the surface with active element (XL-I-177) 101 In Figure 4.5, it can be seen that minimum adsorption distance for mustard gas is around 4.5 ?, while for prenol it is 4.8 ? from PET surface activated with XL-I-177. Interaction energies between gases and active surfaces were analyzed further to investigate the possibility of a second, third and forth layer of adsorption on the surface after it has already attracted and adsorbed the toxic molecules. It was shown that (Figure 4.6), in the case of ammonia, a chemical attraction will occur at around 4.9 ? from the surface. Upon attachment of ammonia on the PET surface, the chemisorbed layer will still be able to physically attract the ammonia molecules and will cause a first layer of physisorption at a distance of around 8 ?, a second layer of physisorption at a distance of 12 ? and a third layer of physisorption at a distance of 15 ? from the surface. Figure 4.6. Interaction energy for ammonia 102 4.4.2. Probabilistic chemi-sorption model A numerical procedure is devised to calculate the passage of the toxic molecules through microporous membranes coated with active elements is calculated based on the following assumptions: a) For every active element and type of gas exposed to an active element, there is a minimum distance between the active site and gas molecule where Ek0 TimeA=Time1*nTotal; %TimeA=Time for Adsorption TimeF=LeftnmolsActual*Time1/nmolscs; %TimeF=Time for flow after saturation TimeTotal=TimeA+TimeF; mol=nmolsActual*kb/RGasConstant; slugx=5; else mol=nmolsPassed*kb/RGasConstant; slugx=nmolsPassed*5/nmolsActual; % slug movement in cm TimeTotal=Time1*nTotal; end %%5. %%%%%%%%%%%%%CALCULATE GTR%%%%%%%%%%%%%%%% Rate=mol/TimeTotal; Rate=Rate/(sheet_sizeactual/1000000); Pinch=PressurePa*0.001*4.014; Timefor5cm=5*TimeTotal/slugx; fprintf(1, '%g \t %3d \t %3d \t %3d \t %g \n', Pinch, Rate, Jk, Jp, Timefor5cm); end 148 C.4. Subroutines C.4.1. Chemisorption function [nTotal, Nc, nmolsPassed, nmolscs]=Chemisorption(thickness, poreD1, nmolscs, nmolsActual, sheet_sizeactual, Na, npores) %Calculates number of molecules adsorbed inside pores, based on Gaussian %assumptions for chemisorption and physisorption: %% Input Parameters: fprintf(1, 'Gas Type= Ammonia \n'); % From Gaussian: dis = 5e-4; % min. chemisorption distance in micron disP1= 8e-4; disP2=12e-4; disP3=15e-4; poreD1t=0; for i=1:npores poreD1t=poreD1t+poreD1(i)^2; end InitialArea=pi*poreD1t/4; AttArea=sum(pi*dis*(poreD1-dis)); %Attraction area for chemisorption Nm=nmolscs*AttArea/InitialArea; %Total # of molecules in chemisorption field in membrane % InitialArea=pi*poreD2t/4; AttArea1=sum(pi*disP1*(poreD1-disP1)); %Attraction area for 1st layer of physisorption NmP(1)=nmolscs*AttArea1/InitialArea; %Total # of molecules in physisorption field in membrane % InitialArea=pi*poreD3t/4; AttArea2=sum(pi*disP2*(poreD1-disP2)); %Attraction area for 2nd layer of physisorption NmP(2)=nmolscs*AttArea2/InitialArea; %Total # of molecules in physisorption field in membrane 149 % InitialArea=pi*poreD4t/4; AttArea3=sum(pi*disP3*(poreD1-disP3)); %Attraction area for 3rd layer of physisorption NmP(3)=nmolscs*AttArea3/InitialArea; %Total # of molecules in physisorption field in membrane %Chemisorption: n=Na/Nm; %Total time instances necessary to fill all active sites Nads=4*Na; %Potential no. of molecules that can be adsorbed by chemisorption and 3 layers of physisorption nmolsPassedTotal=n*nmolscs; %Total no. of molecules that will go through the pore in n instances %Physisorption: for i=1:3 NaLayers(i)=Na; ni(i)=Na/NmP(i); nmolsPassedTotal=nmolsPassedTotal+ni(i)*nmolscs; %Total no. of molecules that will actually pass end nmolsPassed=nmolsPassedTotal-Nads; %number of molecules that passed through the pore until saturation nTotal=n+sum(ni); %Total time instances necessary for saturation C.4.2. Porecount function [poreD, npores, pAve]=porecount(NUMPOREGROUPS, npores, np, pdll, pdul) %Calculates number of pores based on pore groups and pore sizes for each %group obtained from porosity measurements npores=floor(npores); %convert to integers np=floor(np); %convert to integers npk=0; %dum variable for numbering pores pSum=0; %dum variable 150 for k=1:NUMPOREGROUPS for i=1:np(k) j=i+npk; poreD(j)=pdll(k)+(pdul(k)-pdll(k))*rand; pSum=pSum+poreD(j); end %end i loop pAve(k)=pSum/(j-npk); pSum=0; npk=npk+np(k); end %end k loop npores=j; C.4.3. Porosity function [fraction_of_holes]=porosity(npores, poreD, SCALEFACTOR, sheet_size) %%Calculates actual porosity of randomly created pores area_of_pores = 0; for i=1:npores area_of_pores = area_of_pores + (pi*(poreD(i)^2)/4); end %end of npores loop fraction_of_holes = area_of_pores*SCALEFACTOR/(sheet_size); C.4.4. Knudsen function [Jk, Jp]=knudsen(avepD, fraction_of_holes, tp, RGasConstant, Temperature, Vgas, mfp, PressurePa, thickness) %% Calculates fluxes based on Knudsen and Viscous flow equations A=(avepD/3)*(fraction_of_holes/tp)/(RGasConstant*(Temperature+273.15)); %knudsen const B=(pi*(avepD^2)/128)*(fraction_of_holes/tp)/(RGasConstant*(Temperature+273.15)); %viscous Kk=Vgas*A; %knudsen coef 151 Kp=Vgas*B/(mfp/1000000); %viscous coef Jk=Kk*PressurePa/(thickness/1000000); %Knudsen flux Jp=Kp*PressurePa/(thickness/1000000); %Viscous flux C.4.5. Trajectory function [avenewpath]=trajectory(NUMPOREGROUPS, pAve, thickness, mfp) % Calculates the total length of molecular trajectory inside the pore avenewpath = 0; %average of total path length inside pores %%Calculate the length the molecule will travel inside the pore for k=1:NUMPOREGROUPS xlim=pAve(k); %dum variable to pick ave pore diam for group x1=rand*xlim; %2D pore with x as width and z as height z1=0; newpath=0; %another dum variable alfa=rand*180; %direction angle a=0; %number of steps that molecule takes while z190 alfa=180-alfa; xlength=mfp*abs(cos(alfa*pi/180)); if (x1-xlength)<0 length=x1/abs(cos(alfa*pi/180)); x=0; coeff=90; angle=0; 152 else length=mfp; x=x1-xlength; end %ends if (x1-xlength)<0 else xlength=mfp*abs(cos(alfa*pi/180)); if (x1+xlength)>xlim length=(xlim-x1)/abs(cos(alfa*pi/180)); x=xlim; coeff=90; angle=90; else length=mfp; x=x1+xlength; end %ends if (x1+xlength)>xlim end %ends alfa>90 z=z1+length*abs(sin(alfa*pi/180)); newpath=newpath+length; z1=z; x1=x; alfa=angle+rand*coeff; alf(a)=alfa; len(a)=length; end %ends the while loop avenewpath=avenewpath+newpath; end %ends the calculation for molecular trajectory inside pores for all groups avenewpath = avenewpath/NUMPOREGROUPS;