in. {a “.53."... .3“ "map F 4 up... 33“. .1 J. , .3 , . r .; 1.1.11 11...? Ii :3 : . x. .1. 11 3h“ ”8.35.10. .Tilv' m. fix: 0.. Ibis! t!- lxn nun... ‘oixll an: 2- i?! it." 1.15. L! .39. It ( u. .5 5mm? x... ,. .mm. 1.7.9. ‘\ .ll. .2 !NHI.: It 53...”... . 77‘ . ‘3. Au...|:;:|\ 3 .v; .. Ir‘ 7:». A I,\vr (L rt .‘ ‘i. :9. . 2. t 11.11:“; I 1.32... . L a 1 :1. 5i .1. : .3Hr......|:i:ly...'l\ magi. , {awhékafifigfigéyfik é . 4. «a... g... .6. V 43...;5? A THESIS (V Adsorption: 1293 01410 7431 This is to certify that the dissertation entitled A Study Adapting Cubic Equations of State h presented by C * ' Ramkumar Subramanian has been accepted towards fulfillment Ph.D. of the requirements for degree in Chemical Engineering Date /0//0/?5- Majorvprofessor MS U is an Affirmative Anion/Equal Opportunity Institution 0-12771 ‘— LIBRARY Michigan State University PLACE II RETURN BOX to romovo this chockout from your rooord. TO AVOID FINES rotum on or botoro dot. duo. DATE DUE DATE DUE DATE DUE 4%}? ll:— MSU to An Atfirmottvo MbnlEqunl Opporumtty Intuition mam ABSORPTION: A STUDY ADAPTING CUBIC EQUATIONS OF STATE Ramkumar Subramanian A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemical Engineering 1995 ABSTRACT ABSORPTION: A STUDY ADAPTING CUBIC EQUATIONS OF STATE By Ramkumar Subramanian The Simplified Local Density (SLD) method is a new engineering approach to model adsorption based on spatial invariance of the chemical potential along with an equation of state. This work extends previous pure fluid applications to binary mixtures and more complex adsorbent geometries (slits, pores). Model predictions are compared to molecular simulations and experimental data. The SLD model can represent adsorption isotherms of Types I-V at subcritical conditions. Supercritical behavior has not been classified, however the SLD approach can accurately represent the complex behavior exhibited at supercritical conditions. In addition, clustering (molecular charisma) in supercritical fluids is modeled by representing infmite dilution by calculating the fluid density in the region around a single solute molecule. Simple engineering models such as Langmuir and Freundlich cannot represent the variety of experimental adsorption isotherm shapes. On the other hand, molecular simulations represent the behavior, but are not suitable for routine process design. This SLD approach bridges the gap by retaining both the essential physics of the adsorption problem and the simplicity of an equation of state. The SLD approach is demonstrated to be a powerful engineering tool for prediction of adsorption, and requires fewer parameters than previous engineering models. In addition, the SLD approach can be extended to non-ideal gas phase regions without additional parameters. The single parameter used for each adsorbate is temperature- and pressure-independent, permitting extrapolations and predictions of adsorption behavior. Dedicated to my parents Krishnambal Subramanian & K. R. Subramanian, to my sisters Shyamala Nageswaran & Rama Ganesh, and to my late Kadayam grand father iv Acknowledgments First of all, I would like to thank my advisor, Dr. Carl Lira, for his guidance throughout my research, and also for his patience during tough times. I would like to thank Drs. D. J. Miller, R. B. Buxbaum and P. M. Duxbury for serving on my committee and for making appropriate suggestions to guide my research. I would like to thank Dr. Bharath Rangarajan for all the discussions that we have had during this research, and for his valuable suggestions. I would like to thank Arvind and Shalini Mathur for helping me during times of need and for their wonderful company throughout my stay here in East Lansing. I would like to thank my wonderful roommates S. Iyer, and R. Easwar for being the best companions anyone could ask for. I would like to thank Sanjay Yedur for all the wonderful times we spent at coffee shops discussing worldly matters! I would like to thank Mike Bly, Sanjay Padaki, Valerie Adegbite, Satyadev Chilukuri, Kartik Pashupati, Nishant Rao and other unnamed friends for all their support and friendship during my stay at Michigan State. I would like to thank the secretaries, Faith, Julie, Beth and Candy for being so wonderful. I would like to thank my parents Mr. and Mrs. K. R. Subramanian, and my sisters, Shyamala and Rama for supporting me throughout my study. Finally, I would like to thank my wife Jayashree for her love and understanding for the last few months. Michigan State is a great place to live and to study, and for the last six years I have had the greatest time of my life. Thank you Michigan State, and Go Spartans! Table of Contents LIST OF TABLES LIST OF FIGURES NOMENCLATURE (SLD MODEL) CHAPTER 1: INTRODUCTION CHAPTER 2: BACKGROUND, LITERATURE REVIEW AND PRELIMINARY RESULTS DEFINITION CLASSIFICATION OF ABSORPTION ISOTHERMS ADSORPTION MODELS Henry’s Law Langmuir Isotherm BET Isotherm Toth & UNILAN Equations Two Dimensional Equations-of State Potential Theories Polyani Potential Theory Dubinin’s Theory Frenkel-Halsey-Hill Theory van der Waals Model of Barrer and Robins Statistical Mechanical Models Inteng Equation Theory Density Functional Theory Molecular Simulations Mixed-Gas Isotherms Ideal Adsorbed Solution (IAS) Theory PRELIMINARY RESULTS - SIMPLIFIED LOCAL DENSITY MODEL OF RANGARAJAN, LIRA AND SUBRAMANIAN Model Development Outline of Algorithm to Solve for Density Profile Results of SLD Model Ideal Gas - Hard Wall Ideal Gas - Attractive Wall Attractive Fluid - Hard Wall Attractive Fluid - Attractive Wall SLD Model - Discussion vi viii ix xi 35 35 41 45 45 45 46 55 Table of Contents (contd.) CHAPTER 3: QUANTITATIVE MODELING OF ADSORPTION AND EXTENSION OF THE SLD APPROACH TO PREDICT CLUSTERING 58 INTRODUCTION 58 MODEL DEVELOPMENT 60 Algorithm for Solving Density Profile 66 RESULTS 67 DISCUSSION 84 CHAPTER 4: ADSORPTION OF PURE GASES IN SLITS AND PORES, AND ADSORPTION OF BINARY MIXTURES 87 INTRODUCTION 87 MODEL DEVELOPMENT - PURE GAS 89 Slit-like Pores 89 Cylindrical Pores 92 MODEL DEVELOPMENT - BINARY MIXTURES 96 Algorithm to Solve Density Profile 97 RESULTS 98 DISCUSSION 1 l 0 CHAPTER 5: CONCLUSIONS 1 19 APPENDICES 122 REFERENCES 196 vii Table 2-1: Table 3-1: List of Tables Molecular Properties Used in the SLD Model Molecular Properties Used in the SLD Model Tables in Appendix 3 - Properties of Gases Table 3-1 Table 3-2 Table 3-3 Table 3-4 Table 3-5 Table 3-6 Table 3-7 Supercritical Ethylene Supercritical Methane C02 Argon Propane Methane Ethylene viii 47 68 187 189 190 191 192 193 194 Figure 2-1: Figure 2-2: Figure 2-3: Figure 2-4: Figure 2-5: Figure 2-6: Figure 2-7: Figure 2-8: Figure 2-9: Figure 2-10: Figure 2-11: Figure 2-12: Figure 2-13: Figure 2-14: Figure 3-1: Figure 3-2: Figure 3-3: Figure 3-4: Figure 3-5: Figure 3-6: Figure 3-7: List of Figures Adsorption of Ethylene on Graphon -Data of Findenegg (1983) Adsorption of Propane on Graphon - Data of Findenegg (1983) Adsorption of Krypton on Graphon - Data of Findenegg (1983) Variety of Behaviors at Subcritical Conditions Comparison of Molecular Simulation (van Megen and Snook, 1982) to Experimental Data (Specovius and Findenegg, 1978) Grand Canonical Ensemble Monte Carlo Simulation of Ethylene in a Slit (T/Tc = 0.85) Grand Canonical Ensemble Monte Carlo Simulation of Ethylene in a Slit (T/T, = 1.2) Density Functional Theory - Comparison to Grand Canonical Ensemble Monte Carlo Simulation F ugacity Vs Pressure for a Pure Fluid Adsorption of Ethylene on Graphon Adsorption of Ethylene on Graphon Density Profile of Ethylene on Graphon at 283.15 K Weak Adsorption of Ethylene on Graphon Adsorption of Krypton on Graphon Adsorption of Ethylene on Graphon Adsorption of Krypton on Graphon Adsorption of Propane on Graphon Adsorption of Argon on Graphon Adsorption of Methane on Graphon Density and Excess Number of Solvent Molecules within a Sphere of Radius R around a Solute Number of Excess Solvent Molecules within a Sphere of Radius L around a Solute 10 28 3O 31 32 43 49 50 51 53 54 69 71 72 73 74 76 77 Figure 3-8: Figure 3-9: Figure 3-10: Figure 3-11: Figure 3-12: Figure 4-1: Figure 4-2: Figure 4-3: Figure 4-4: Figure 4-5: Figure 4-6: Figure 4-7: Figure 4-8: Figure 4-9: Figure 4-10: Figure 4-11: Figure 4-12: Figure 4-13 Figure 4-14 Figure 4-15: Clustering of CO2 around Naphthalene Clustering of CO2 around Naphthalene Comparison of Cluster Size with Fluorescence Spectroscopy - CO2 around Naphthalene Comparison of Cluster Size with Fluorescence Spectroscopy - CO2 around Pyrene Comparison of Cluster Size with Fluorescence Spectroscopy - Ethylene around Pyrene Slit-Like Pores Cylindrical Pores Grand Canonical Ensemble Monte Carlo Simulation of Ethylene in a Slit (T/T, = 0.85) Ethylene on Carbon Slits at T/Tc of 0.85 Ethylene on BPL Carbon at 212.7 K Ethylene on BPL Carbon at 260.2 K Ethylene on BPL Carbon at 301.4 K Methane on Activated Carbon at 243.15 K Adsorption of Ethylene-Methane Mixture on BPL Carbon at 212.7 K and Initial Bulk Ethylene Concentration of 0.74 Adsorption of Ethylene-Methane Mixture on BPL Carbon at 212.7 K and Initial Bulk Ethylene Concentration of 0.74 Adsorption of Ethylene-Methane Mixture on BPL Carbon at 260.2 K and Initial Bulk Ethylene Concentration of 0.235 Adsorption of Propylene-Ethylene Mixture on BPL Carbon at 293.15 K and 1.013 bar a(r)/ab for a Pore that can hold 2.5 Ethylene Molecules in the Horizontal Plane Adsorption of Homogeneous Ethylene on Graphon Adsorption of Homogeneous Ethylene on BPL Carbon at 212.7 K Figure Appendix 3-1 Error in the Peng-Robinson Calculation of Saturation Density 78 79 80 81 82 90 93 99 100 102 103 104 106 107 108 109 111 113 ' 115 116 195 Nomenclature (SLD Model) a, b Constants of the van der Waals or Peng-Robinson equation f fugacity g(r) radial distribution function N A Avagadro’s number pressure T temperature 2 distance from surface of the wall. Intermolecular distance is z+ofS/2 Z compressibility factor k Boltzmann’s constant v molar volume nex number of molecules A s Surface Area Greek 1 F“ surface excess 4) fluid-fluid pair potential [.1 chemical potential ‘1’ fluid-wall potential (physical adsorption), fluid (solvent)-solute potential (clustering) p molar density a molecular diameter a pair interaction parameter xi Subscripts at! attractive property rep repulsive property jf contribution due to fluid-fluid interaction f fluid property 3 solid or solute property (in clustering) fir contribution due to fluid-solid interaction (physical adsorption) or fluid (solvent)-solute interaction (clustering) b bulk property (uniform fluid) 0 standard state xii CHAPTER 1: INTRODUCTION Physical adsorption is a separation process in which certain components of a fluid phase are transferred on to the surface of a solid adsorbent [McCabe, Smith & Harriott, 1984]. It is the result of intermolecular forces of attraction between the fluid and solid molecules. If the attractive force of the solid on the fluid is greater than the intermolecular forces among the fluid themselves, the fluid will condense upon the surface of the adsorbent [T reybal, 1983],and may be accompanied by the evolution of heat. In engineering applications, adsorption is usually used an alternate to extraction. Adsorption tends to have a smaller capacity but a higher selectivity than extraction. It can also be more gentle for e.g. proteins may be denatured by organic extraction but may be adsorbed. Separation processes using adsorption are on the increase while those using extraction are on the decrease [Belter, Cussler & Wu, 1988]. Physical adsorption is used in gas purification processes such as the removal of volatile organic compounds from stack gases, as a means of fractionating fluids that are difficult to separate by other methods, and in adsorbent regenerations using supercritical fluids [Tan and Lion, 1988, 1990]. Physical adsorption is also of interest in transportation and storage of fuel and radioactive gases, separation and purification of lower hydrocarbons, solid-phase extractions fluids, in supercritical extractions and chromatography [Findenegg, 1983; Barton, 1983; Strubinger and Parcher, 1989], and in critical point drying [Rangarajan and Lira, 1992]. Understanding the thermodynamics and structure of the gas- solid interface is essential to the understanding of heterogeneous catalysis and wetting phenomena [Findenegg, 1983]. While there are a large number of theoretical and experimental studies of adsorption below the critical temperature, there are few studies of adsorption near or above the critical temperature of the fluid. Such studies are especially relevant to the storage of methane at ambient temperatures (T, = 1.57) at reasonably high densities. Theoretical approaches to understanding and predicting adsorption range from simple empirical fits (Freundlich/Tom isotherms) to theoretically-sound methods such as molecular dynamics (MD) and Monte-Carlo (MC) simulations. Computer simulations such as the grand canonical ensemble Monte-Carlo semi-quantitatively predict the cusp- like behavior near the critical point [van Megan and Snook, 1982]. However, such methods are computationally intensive. Simulations are difficult near the critical point due to fluctuations, and require a large number of molecules and consequently significant amounts of supercomputer time. Statistical mechanical theories such as the density functional theory are also computationally intensive although they are about two orders of magnitude faster than computer simulations [Gubbins, 1990]. On the other hand, the traditional empirical and semi-empirical methods which are computationally undemanding are unable to account for the wide variety of shapes of adsorption isotherms seen near the critical region. There is an engineering need for a model that bridges the gap between the simple and sophisticated models, and has predictive capability. Further, the development of processes utilizing supercritical fluids requires engineering models capable of spanning large pressure ranges. Experiments with supercritical fluids are more difficult and expensive than experiments at atmospheric conditions, and a model that can predict supercritical behavior would be a well-received developmental tool. The focus of this thesis is to develop the simplified local density (SLD) model, and demonstrate its application in predicting pure fluid adsorption isotherms over wide pressure and temperature ranges by comparing model predictions with experimental data from literature. The SLD model developed in this work superimposes the fluid-solid interaction potential on the van der Waals and Peng-Robinson equations—of-state. My intention is to present the concepts necessary to adapt common cubic equations of state for describing the adsorption phenomenon. The SLD model is intended to bridge the gap between the computationally intensive but more theoretically sound statistical mechanical models and the undemanding, empirical methods. The SLD concept should be a useful engineering supplement to other available models, and is not intended as a replacement for the statistical mechanical models. With the above objectives, the thesis has been divided into the following chapters: 1) Introduction 2) Background, Literature Review and Preliminary Results. 3) Quantitative Modeling of Adsorption and Extension of the SLD Approach to Predict Clustering 4) Adsorption of Pure Gases in Slits and Pores, and Adsorption of Binary Mixtures 5) Conclusions and Recommendations In chapter 2, a detailed literature survey of the various theories of adsorption is given, followed by the development of the van der Waals SLD model. High pressure experimental adsorption data that show fascinating behavior are also given in this chapter. The basic assumptions leading to the SLD model, and the model equations are described in this chapter.. In chapter 3, the van der Waals equation is replaced by the Peng-Robinson equation-of-state, thus making the model a quantitative predictive model. Model results for adsorption on flat walls are compared to experimental data. The model is modified to describe clustering in supercritical fluids, and model predictions are compared to experimental fluorescence spectra. The model is extended to describe adsorption in slits and cylindrical pores in chapter 4. Model predictions for adsorption of pure fluids are compared to both experimental data and molecular simulations. The model is extended to describe adsorption of binary mixtures in slits. Model predictions of both the concentration and amount adsorbed are compared with experimental data. Finally, the results of this model are summarized in chapter 5, followed by recommendations for future work. CHAPTER 2: BACKGROUND, LITERATURE REVIEW AND PRELIMINARY RESULTS DEFINITION While several different methods are used to measure adsorption, adsorption is un- ambiguously defined by the surface excess, the excess moles per unit area. It A S (X I‘CX = L:[p(z)-p.]dz [2-1] where p (z) is the density at a distance 2 from the wall, and pb the bulk density. 2,, is often defined as the plane above the surface of the solid where the fluid-solid potential is zero. Before reviewing the theories for adsorption, some experimental results will be presented. Reliable experimental measurements of adsorption of fluids on well character- ized solids are essential in order to test different theories of adsorption. While most com- mercial adsorbents are porous and heterogeneous, graphitized carbon black (g.c.b.) or exfoliated graphite or the basal planes of graphite provides an atomically smooth homo- geneous and non-porous surface to study adsorption. Graphitized carbon black (or Gra- phon) provides the ideal surface to understand physical adsorption and wetting, and theo- ries for adsorption on homogeneous surfaces can be extended to heterogeneous and po- rous materials. Some of the best data available are those of Findenegg and co-workers who have examined the adsorption of a variety of gases such as ethylene, propane, meth- 5 6 ane, krypton and argon on graphon [Findenegg, 1984; Specovius and Findenegg, 1978, 1980]. Some of their results are presented in Figures 2.1 - 2.3, where one sees the ad- sorption isotherms of subcritical propane and ethylene, near-critical propane and ethylene, and supercritical ethylene and krypton, on graphon. CLASSIFICATION OF ADSORPTION ISOTHERMS Adsorption isotherms (subcritical) are classified into six types [Sing et al., 1982; Sing 1983], as shown in Figure 2.4. Type I isotherms are normally seen in microporous adsorbents, Type IV & V on mesoporous adsorbents and Type II, III and VI on non- porous adsorbents. Types III & V are similar to Types II & IV respectively except for the stronger magnitude of the adsorbate-adsorbent interaction in the latter cases. Type VI isotherms indicate layering, and are observed for the adsorption of Ar/Kr on g.c.b. at low (liquid N2) temperatures [Bienfait 1980 and references therein]. ADSORPTION MODELS In this section some of the basic theories for physical adsorption will be briefly re- viewed. There is an enormous amount of literature available on adsorption and the reader is referred to several monographs [Young and Crowell, 1962; Brunauer, 1945; de Boer, 1953; Defay and Prigogine, 1966; Nicholson and Parsonage, 1982; Yang, 1987; Lee, 1988]. a 43—26315 K 40 .. +273.15 K +283.15 K 35 __ +293.15K I +313.15 K a" i g 30 .. 8 3: 25 -- III e 20 -- 0 X I.” I5 - IO - 5 - 0 u I I I I I 0 20 4O 60 80 100 120 Pressure (bar) Figure 2-1: Adsorption of Ethylene on Graphon - Data of Findenegg (1983) 30 -- 25 -- A N 20 -- g o S. .v 15 -- In in o o x In 10 -- _. . . - __ ~ +273.15K 5 ; .‘ ~—~ ‘ +323.15K (1,7,; " +343.15K .. if +363.15 K 0 ' I : : I : : :' 0 5 IO ‘15 20 25 30 35 Pressure (bar) Figure 2-2: Adsorption of Propane on Graphon - Data of Findenegg (1983) 40 14 12 -- IO ‘- ‘MA :2 O 8 .. .. E :3. V 8 6 -- O O X Ill 4 -_ +253.15K +273.15K +298.15K 2 -_ +323.15K I +348.15K "I +373.15K 0 I I I 0 4O 80 120 Pressure (bar) . Figure 2-3: Adsorption of Krypton on Graphon - Data of Findenegg (1983) WomwQanmhv/N «Clogs/N Fig j .0 (D e. O U) .0 <3: '5 3 O E < Y In P / P58“ Figure 2-4: Variety of Behaviors at Subcritical Conditions 11 Henry's Law In the low relative pressure region, the adsorption is linear, and the amount ad- sorbed (n) is proportional to the pressure (P). This region is often called the Henry's law region. n = k P [2'2] where k is the Henry’s law constant. Sometimes the pressure at which this law is obeyed is below experimentally observable pressures. If the adsorption is measured at sufficiently low pressures discontinuities in the adsorption-pressure curve are sometimes noticed indi- cating phase transitions in the adsorbed phase as shown by the results of Ross and Clark [1954], Fisher and McMillan [1957], and more recently by a host of modern techniques by Thomy and Duvall [Bienfait 1980]. Langmuir Isotherm One of the most popular empirically-adjusted models for adsorption is the Lang- muir isotherm, a two parameter equation. This is one of the most popular isotherm due to its extreme simplicity, and is often expressed as e = = [2-3] 12 where W is the mass adsorbed, and W... the amount adsorbed in a monolayer. The Lang- muir equation is only derived for monolayer adsorption and therefore predicts only Type I isotherms. However, it is frequently applied to microporous materials. It is more suitable for describing chemisorption. BET Isotherm The Brunauer, Emmett and Teller (BET) theory extends the Langmuir theory to multilayers. It assumes that molecules adsorb in stacks or layers and that the uppermost molecules in the adsorbed stacks are in equilibrium with the vapor. The BET theory treats the first adsorbed layer separately and assumes that all layers above the first layer are con- densed, and treats them identically as a liquid . This theory predicts that I’ll”, C P P P (“EX“TCE) 9-1-1- [2...] Wm NM where P, is the vapor pressure, N the number of moles adsorbed and C a constant related to the fraction of surface that is uncovered. The BET equation is often used in deterrnin- ing pore-size distributions and is often expressed as 1 1 C—lfi W01; _1) = CW," + CW", P0 [2.5] Tl 13 In pores where the number of layers are limited to n the BET equation is : C{l—(n+1)[—I:] +n[£] } N _ P. P. N — n+1 .. (£_1){1+(C—1)P_C[_P_) } P P0 P0 W e _ W. _ [2-6] This equation reduces to the previous equation when n is infinity and the Langmuir isotherm for n = 1. The BET theory and its modifications are very p0pular since they Show all the different types of subcritical isotherms. However, some of its limitations are due to the assumptions made in the developing the theory 1) the surface is energetically homogeneous; 2) there are no lateral adsorbate interactions; 3) adsorption is localized - adsorbed molecules are fixed on sites and do not laterally move; 4) the heat of adsorption in the second and higher layers is the same as the heat of condensation. Most of these as- sumptions are valid for pressures in the range 0.05 < P/P‘ < 0.35, where the high energy sites are filled but extensive multi-layer condensation has not commenced. The BET the- ory is limited to subcritical temperatures, and assumes an ideal gas vapor phase, although corrections can be applied. Toth & UNILAN Equations The Toth equation can be written as l4 mP n = ———l',7 [2-7] (b + P’) This is a three parameter equation m, b, and t, where m is the saturation capacity (mol/kg), b and tare constants of the Toth equation, n is the specific amount adsorbed, and P is the pressure. A similar equation called the UNILAN equation can be written n = 31n[c+ Pi] [2-8] c+Pe 23 where sis a constant. Both these equations reduce to the Langmuir equation (Toth - t = l, UNILAN - s = 0). Both these equations are empirical equations, but are for heteroge- neous adsorbents. The constants needed are temperature-dependent and atleast two iso- therms must be measured experimentally before calculating the temperature coefficient of adsorption. However, these theories seem to be the best overall for microporous adsorb— ents, and have been used as the basis for multicomponent adsorption models which will be discussed later in this chapter. Two Dimensional Equations-of-State The BET theory of multi-molecular adsorption is a generalization of Langmuir's unimolecular theory. Both theories are based on localized physisorption, and both neglect lateral interactions between fluid molecules, i.e. interactions between molecules in the Same layer. A different approach treats the adsorbed phase as a two-dimensional fluid, 15 and uses a two-dimensional equation-of-state to represent the adsorbed layer [Hill, 1946, 1947, 1948; de Boer, 1953]. The two-dimensional equations-of-state allow for interac- tions between fluid molecules in the first adsorbed layer. Neither the Langmuir nor the BET theory allow for interactions between molecules, either on the surface, or between molecules in different layers. When a fluid is adsorbed onto a surface it exerts a pressure opposing the surface tension of a clean surface. If yo is the surface tension of the clean surface and 7 that of the surface with an adsorbed fluid, then the spreading pressure It is defined as ”=70“? [2-9] The surface tension 7, or 1: can be related to the Gibb's surface excess 1‘" = "1391 [2.101 where a is the activity. For a dilute solution, the activity (a) may be replaced by the con- centration C or pressure P, and a two-dimensional equation of state is used to give 1:. For an ideal gas the equation is mlzRT [2-11] 16 where A is the area per mole. This equation implies no forces or interactions between the fluid molecules, which is what one would expect at low coverage. These two equations directly lead to the Henry's law isotherm. Such an approach along with activity coeffi- cients has been used to model adsorption from gas mixtures and from solutions [Myers and Prausnitz 1965; Radke and Prausnitz 1972]. Several attempts have been made using a two-dimensional equation-of-state to represent the adsorbed layer which accounts for mobile interactions and includes lateral interactions in the first layer (The Langmuir and BET models account only for localized adsorption - i.e. they neglect lateral interactions between fluid molecules) [I-Iill, 1946, 1947, 1948; de Boer, 1953]. The two dimensional van der Waals equation predicts a two- dimensional critical temperature Tc; which is one half the bulk critical temperature (T c2 = 0.5 T C) and a two-dimensional critical pressure (I), = 0.361 P, ( Vc / N )1”. Here the sub- script c2 is used to describe the critical constants for the two dimensional state. Similar results for Tc; are obtained with different systems. A Lennard-Jones Devonshire [1937] model gives T c2 = 0.53 Tc. An excellent discussion of two dimensional critical tempera- tures and pressures from the van der Waals equation-of-state is given by de Boer [1953]. Typically Tczl T c is about 0.4 with extremes of 0.36 and 0.56. This lies in-between mean field theory prediction of 0.5 and a 2-D Ising model prediction of 0.37 [Bienfait 1980]. The two-dimensional van der Waals model leads to the following equation O e]exp(-k, G) [2-12] P —=kOex P . ,[1 0 17 The constant k, depends on a2 and b2 the two-dimensional van der Waals parameters, and therefore only the adsorbate. A high value of k. implies strong inter-molecular forces (fluid-fluid). The constant k; measures the strength of adsorption, and the stronger the adsorption the smaller the k2. This equation predicts phase transitions on the surface for weak adsorption forces and strong intermolecular attraction (i.e. large k, and k2). If the fluid-fluid intermolecular attraction is weak, no condensation is seen. For strong adsorption forces, a tendency to- wards saturation (Type I isotherm) is seen, and the adsorbed fluid is either a compressed supercritical fluid or a condensed two dimensional liquid. Relative to the BET equation, the Hill de-Boer model predicts lower pressures for coverage's below 0.8, but over-corrects the errors of the BET equation. This is due to the fact that lateral interactions are allowed, and not due to the mobility. If lateral interactions are allowed in a localized first layer, condensation takes place at very low pressures, and jumps from almost zero to near unity at a critical coverage 9, of 0.5. Such a result could probably be deduced by percolation on a lattice, since the similarity between the abrupt jump at 9c of 0.5 and a critical percolation probability of pc of 0.5 does not seem coinci— dental. 18 Potential theories Several other popular theories are the Polyani potential theory [Polyani 1932] and modifications of it such as the Dubinin-Radushkevich [Dubinin 1947, 1975], and the Frenkel-Halsey-Hill [Halsey 1948; Hill 1950] theory. (i) Polyani Potential Theory The Polyani potential theory assumes that a potential field exists at the surface of a solid that attracts the molecules of the surrounding fluid. These attractive forces are es- sentially London dispersion forces or van der Waals forces, which vary as 1/r3 (where r is the distance from the surface). The adsorption potential (8;), at a distance i from the surface is the work done in bringing a molecule from the gas phase at density Px to that point where the density is p i° Polyani noticed the similarity with gravitational potential, and used the equation of hydrostatics to calculate the adsorption. e, =[p'vdp [2-13] I It Was assumed that the adsorbent potential at a given distance (or surface at) i is e,- , re- gardless of the number and kind of molecules located between the i'th surface and the SOlid or gas, an assumption which was justified later by London's theory of van der Waals fOl'ces. Each of the potential surfaces along with the surface of the solid adsorbent en- clOSes a volume 6;. At the surface of the adsorbent 80 is maximum and (l),- is zero. The 19 gas phase is assumed to be the region in which 8 = 0, and limits the adsorption volume to d) = ‘l’max- In actuality, the potential never goes to zero but for practical purposes 8 = 0 at a finite distance and therefore a finite volume from the surface. The theory assumes that de/dT = 0 implying that do/dT = 0 and therefore that e = f (4) ) for all temperatures for a given adsorbent. This is used to form a characteristic curve, which is experimentally de- termined. The amount adsorbed (n) is given by 41.... n= [0 (p—p.)d¢ [2-14] In using the theory, first a characteristic curve 8 = f (q) ) is created. This curve should span the entire adsorption space, and is therefore typically evaluated at a moderate tem- perature. The second step consists of evaluating adsorption isotherms from these charac- teristic curves. The characteristic curve equations can be simplified if the temperature is significantly below or above the critical temperature. These equations ignore adsorption in vapor form, and the compressibility of the liquid layer, as well as assume ideal behavior. Near Tc these assumptions result in errors, but the different errors fortuitously cancel re- sulting in fairly accurate predictions. For high temperatures and pressures the Polyani theory has been modified by using empirical correction factors to account for the com- pressibility and thermal expansion of the adsorbed liquid [Berenyi 1923; Brunauer 1945]. 20 (ii) Dubinin's Theory An empirical relationship was found between 80 of Polyani theory and the van der Waals constant 'a' of the fluid : 5,, = NE [215] which originates from the mixing rules of Berthelot, who proposed that the attractive po- tential between two molecules at a fixed distance is proportional to ,Ia,a2 where a 1 and a; are their van der Waals constants. Using this, the potential theory was extended by Du- binin and co-workers to include adsorption of mixed gases onto a solid. They expressed the characteristic curve for all gases on the same adsorbent as : 8 = Bf(¢) [2-16] where B is called the coefficient of affinity and is defined as the ratio of adsorption poten- tials (B = 8 ISO) of a gas and a reference fluid, on the given adsorbent. They found that the molar volumes of the adsorbate in liquid state were closely proportional to the coefficient of affinity. The characteristic curves resembled the positive side of a Gaussian curve which lead Dubinin and Radushkevich to assume that the adsorption volume V occupied by the liquid adsorbate (and V, by the reference) followed V = V0 exp[—K[%) ] - [247} where K is a constant depending on the shape and size distribution of pores. This leads to 2 2 RT P0 ln( W) — ln(VoP) - K17?) [ln(?)] [2-18] Subcritical vapor pressure data have been extrapolated into the supercritical region and a good fit has been obtained using the extrapolated fugacities [Maslan et al., 1953]. (iii) Frenkel-Halsey-Hill Theory This theory presumes that the adsorbed film consists of a liquid at a density pl (density of bulk liquid at that temperature) with a thickness of h. The surface excess (fer) is given by I‘" = h(pI —p) [2-19] Where p is the density of the bulk gas. Assuming that the adsorbate and adsorbent interact tlll‘ough a 12-6 Lennard-Jones potential (and integrating to a 9-3 potential), this leads to ln(-5] = —-k— [2-20] where k is a constant. In a further modification of the Frenkel—Halsey-Hill theory, multiple liquid like slabs and fugacities have been used to represent the adsorbed layer that has a density p... for 2 S 2. and p; for z, < z < 2:, where the fugacity is related to adsorption by : “’35 5"— [2-21] kT 2,1 5‘ m b» Ix w I: Here f, is the fugacity of the saturated vapor, z, the distance of the i'th statistical layer of molecules, AI: is the perturbation energy of a molecule in the first layer at a distance 21. For a single step density profile if the density of the layer is p; for z < z,- and p for z > 2,, then 2,. can be determined from experiments and [2-22] and the parameter m fit from a plot of 1n(z,-) vs. ln(-ln(fo/f)). Such methods with multiple- step density profiles have been used to describe adsorption generally in the sub-critical range [Findenegg 1983]. 23 van der Waals model of Barrer and Robins This theory, developed by Barrer and Robins [1951], has been unrecognized for more than forty years. It uses the fact that the chemical potential is constant at all dis- tances from a wall, and can be determined by the temperature and pressure far from the surface. This potential consists of two components - potential due to the interaction of the fluid molecules themselves, and a potential due to interaction of the fluid and the wall. Since the potential exerted by the wall and the bulk fluid potential are known the density profile and surface excess can be calculated. They assumed that the fluid obeyed the van der Waals equation-of-state and the fluid-wall interaction can be approximated by the 9-3 Lennard-Jones potential. This leads to the following equation 0 exp( 0 _2a0_ SIN )= 9,, exp 0,, _2a0b [2_23] 1-9 1-9 bRT rRT l-Ob l-Bb bRT where a and b are the van der Waals constants, N and n are the total number of molecules and moles in the system respectively, 0 = bn/V, where V is the total volume occupied by closely packed molecules, r is the distance between the fluid molecule and the surface, C is an empirical constant and the subscript b stands for bulk properties. If v is the molar vol- ume of the fluid being adsorbed on the surface of the adsorbent then equation 2-23 may be written as 24 l b 20 CN 1 b 2a exp - — 3 = exp — [2-24] v—b v—b vRT rRT vh—b vh—b vhRT This model predicted the different adsorbent isotherms whose magnitude could be manipulated by changing the empirical constant C. Hill [1952] suggested that this model can be improved by treating the fluid not as a bulk fluid as Barrer and Robins had done, but as an inhomogeneous fluid i.e. the chemical potential should not be calculated from the local density but from the entire density profile. He also suggested replacing the inaccu- rate van der Waals equation by a better equation. Statistical Mechanical Models There are different types of models derived from chemical physics, including the Lattice Gas model, Integral Equation theory, and Density Functional theory. The Lattice Gas model is a case of localized adsorption, where the molecules of the solid are large compared to the molecules of the fluid. This results in the formation of a lattice of ad- sorption sites that are separated from each other by high potential barriers. This model simplifies to the Langmuir or BET approaches and predicts several phase transitions on the surface. Theoretical interpretations of adsorption include virial expansions in terms of bulk density at low pressures [Steele 1962]. A van der Waals theory for adsorption has been proposed by Sullivan [1979]. This approach has been generalized to examine wetting transitions at the solid-gas interface [Tarazona and Evans 1983], adsorption and wetting 25 transitions in binary fluid mixtures by Telo de Gama and Evans [1983]. The van der Waals density functional model for inhomogeneous fluids has also been used to study wetting transitions at surfaces [Kung et al., 1990; Teletzke et al., 1982a, b]. More recent applications of statistical mechanics have involved adaptation of integral equations devel- oped for homogeneous fluids to inhomogeneous adsorbed fluids. These include integral equations for liquid state theories such as the BBGKY hierarchy for non-uniform liquids [Lee, 1987; Henderson, 1992] and density functional theory [Saam and Ebner, 1978; Fisher and Methfessel, 1980; Tan and Gubbins, 1990; Kierlik and Rosinberg, 1992]. (i) Integral Equation Theory Statistical mechanical theories use distribution functions to describe the correlation between fluid molecules and hence the short range order. The primary cause of this short range order, or structure, of a uniform fluid are the repulsive forces, with attractive forces forming the background. In order to get a complete theory of liquid state an expression is needed for g( r) the radial distribution function [McQuarrie, 1976; Rowlinson and Widom, 1982], or the pair correlation function. -pu 2 I e ”dr...dr V N. i I 3 N [245] gm = N2(N-2)! 2,, Where, there is a system of N particles in a volume V and at temperature T, ZN is the con- flgur‘ational integral, r1, r2, r3 are coordinates in space. If the total potential of a N-body 26 system can be assumed to be pairwise additive, then all the thermodynamic properties of the system can be derived from the radial distribution function. For e. g. the pressure is related to g( r) by 3:— "Pi” ' 4 2d [226] kT—p 6kT 0 ru(r)g(r) Itr r - A number of approximate equations have been used to solve Equation 2-26. These in- clude the BBGKY hierarchy and Omstein-Zernike equations with the Percus-Yevick and hypernetted-chain closures. These approaches are reviewed by McQuarrie [1976] and Henderson [1992], and are widely used in predicting clustering in supercritical fluids. (ii) Densig Functional Theog The density functional approach was derived by Saam and Ebner [1978]. It is based on the fact that there exists a functional 9 of the local density p( r), such that at equilibrium this functional is minimized. F = F[f(r)s p(r), C(r‘ ,r2 )] [2'27] Where f( r) is the local free energy, p( r) is the local density and c( r1, r2) is the direct corre- lation function. Gubbins and co-workers have used this approach to describe adsorption is slits and cylindrical pores and have successfully compared their results to molecular simu- 27 lations. Kierlik and Rosinberg [1991, 1992] proposed a simplified version of the free en- ergy functional for the inhomogeneous hard-sphere fluid mixture that requires four distinct weight functions and generates a triplet direct correlation function for the one component fluid. Molecular Simulations While there a large number of complicated theories they are not exact. The best solutions, for a given potential, are obtained by computer simulations where the equations of motion are solved for the particles of an ensemble. Observable macroscopic quantities are then calculated by time-averaging the appropriate microscopic equations. Such a cal- culation is called a molecular dynamics (MD) calculation. In the Monte-Carlo (MC) scheme a large number of configurations are sampled, and particles moved to find the most stable state, and averages are calculated to give a certain macroscopic property. Computer simulations are difficult to perform near the critical point due to the large density fluctuations. However, these techniques are very powerful, and yield the best results for a given potential. van Megan and Snook [1982, 1985] have simulated the adsorption of fluids near the critical region using a grand canonical ensemble Monte-Carlo method. They compare the results of their simulation with experimentally measured [Specovius and Findenegg, 1978] isotherms of ethylene on graphite, just above the critical temperature and find reasonably good qualitative agreement with data (Figure 2.5). The simulation results show cusp-like maxima which are close in magnitude to experimental values, but the location of the cusp is off by 25% (~15 bar). Their results cannot be 28 20 ‘ L0 50 80 ‘ 100 plborl - Figure 2-5: Comparison of Molecular Simulation (van Megen and Snook, 1982) to Experimental Data (Specovius and Findenegg, 1978) '- - Molecular Simulation (T / r, = 1.03) -- Experimental Data (T / Tc = 1.02) - Experimental Data (T / T‘: = 1.04) 29 quantitatively compared to experimental data because their simulation of bulk ethylene uses a shifted potential and does not represent the experimental bulk properties of ethyl- ene; e.g. the simulation critical temperature (Tc, ,,-,,,) of ethylene is 226 K, as opposed to an experimental value of 282.4 K. However, their results have been the basis for comparison for the density functional theory using the same shifted potential. Figure 2.6 and 2.7 Show the adsorption of ethylene on a slit like pore at sub and supercritical temperatures. At low slit widths Type I isotherms are seen. There is also a slight decrease in the surface excess with increasing pressure. As the slit width increases type IV isotherms are seen. In flat walls, type H isotherms are seen. Figure 2.8 shows the density functional theory’s predic- tion of van Megen and Snook’s simulation [Tan and Gubbins, 1990], and it can be seen that the density functional theory does a very good job of reproducing the results of the molecular simulation. Snook and van Megen conclude that traditional adsorption theories will not be able to account for the wide variety of shapes of adsorption isotherms encoun- tered in their computations. Mixed-Gas Isotherms Current models that predict mixed-gas isotherms use pure gas models such as the Toth or UNILAN equations. One such theory is the ideal-adsorbed-solution theory of Myers and Prausnitz [1965]. This theory is thermodynamically consistent and exact at the limit of zero pressure. The success of such theories depends of their ability to predict the pure gas isotherms accurately. Small errors in the prediction of pure-gas experimental data may cause large deviations in the mixed-gas predictions. Other sources may be the 30 0 '2 ’4 -6 -8 ,1-0 Figure 2-6: Grand Canonical Ensemble Monte Carlo Simulation of Ethylene in a Slit (TIT c=0.85) -h=~, - - - 11:10.0, ----h=5.0, - ----- h=3.5 31 O .1 «2 .. D Figure 2-7: Grand Canonical Ensemble Monte Carlo Simulation of Ethylene in a Slit (T IT e=1 .2) -h=~,~uh=50,~~uh=35 . 32 OUIIIISIIIIILI'4':IIII 0 0.1 0.2 0.3 0.4 . p Figure 2-8: Density Functional Theory - Comparison to Grand Canonical Ensemble Monte Carlo Simulation 33 neglect of surface heterogeneity and adsorbate-adsorbent interactions, both of which may cause non-ideal behavior [Valenzuela and Myers, 1989]. Later models of this type, did not neglect these heterogeneities and were called non-ideal adsorbed solution theories. Since the basis of mixed-gas isotherm modeling is the IAS theory, we shall discuss it in more detail. Ideal Adsorbed Solution (IAS) Theogz For a multicomponent system containing N adsorbates, the IAS equations for a perfect gas are Py.=P."x. (i=1,2, ..... ,N) [2-28] where the adsorption isotherm of pure ith adsorbate is n,"(P,-"), y,- is the mole fraction in the gas phase and x,- is the mole fraction in the adsorbed phase. ‘I’,°(P,") = ‘11: (P20) = ...... = \I';(P,3) [249] ix, =1 [2-30] The function ‘I’,°(P,~") is the integral for the spreading pressure (ILA/RT) for the adsorption 0f Pure ith adsorbate at the same temperature as the mixture. 34 M_Pn _»wmm ——_[ FdP—J d [2.311 n RT 0 0 d ln(n) The independent variables are T, P, y1,...y~.r, leaving us with 2N unknowns (Pf, x.) and 2N equations (2-28 - 2-30). After these three equations are solved, the total amount n, and the amount of ith component are found as l N x. — = 2-3— [2—32] n: i=1 "i n.- = M.- [2-33] The general algorithm for solving this theory is given in Valenzuela and Myers [1 989]. Further development of non-ideal adsorbed solution theories, and the develop- ment of the fast-IAS theory, where the integrals for the spreading pressure are written as a series expansion in terms of the central moments of the adsorption energy distribution, are also given in Valenzuela and Myers [1989]. All these theories require the pure gas iso- therm parameters (for Toth/UNILAN it is 3), and hence are not predictive models. 35 PRELIMINARY RESULTS - SIMPLIFIED LOCAL DENSITY MODEL OE RANGARAJAN, LIRA AND SUBRAMANIAN This model, originally developed by Rangarajan [1992], and Rangarajan, Lira and Subramanian [1995] forms the basis for the rest of this dissertation, and hence will be dis- cussed in detail. Some of the predictions of this model will also be presented and com- pared with experimental data. The model has a basis similar to the Polyani potential the- ory, where the adsorbent exerts an attractive potential on the adsorbate. The model is also Similar to the approaches of Barrer and Robins [1951], Hill [1949], Sullivan [1979], and Kung et al. [1990]. Model Development The attractive potential between the fluid and solid, at any point z, is assumed to be independent of temperature and the number of molecules at and around that point. At equilibrium, the molar chemical potential [.1 must be uniform throughout the system [Denbigh 1981; Guggenhiem 1967] and is calculated by contributions from fluid-fluid and fluid-solid interactions. u = u, = 11,12) + 11.12) [2-341 where the subscript 'b' refers to the bulk, the subscript 'fj“ to the fluid-fluid interactions and the subscript 'fs' to the fluid-wall interactions. Equation 2-34 requires that at a distance 2 36 from the wall, the sum of the chemical potential due to the fluid-fluid interactions and the attractive potential exerted by the solid on the fluid remains constant, and equal to the bulk chemical potential. While such an equation has been written from the principles of Cherrlical equilibrium, this fundamental result for inhomogeneous systems has been derived by minimizing the grand potential [Evans 1979; Rowlinson and Widom 1982; Davis and Scriven 1982]. In using this equation care should be taken to ensure a consistent basis for chemical potential, whether it be per molecule or mole. If ‘I’( z ) is the potential exerted by the wall on a fluid molecule , then on a molar basis, 11,, (z) = N.‘P 0, b > 0, and ‘I’(z) = 0, f(z) = fb. The InOde] predicts a reduced density near the wall, because of attraction of the fluid molecules near the wall to fluid molecules in the bulk. In fact a vapor-liquid phase transition is seen 46 adjacent to the wall for certain cases when the bulk fluid is liquid. This is similar to the observations of Abraham and Singh [1978]. 4. Atflctive Fluid - Attrzyltive Wall For this case, a > 0, b > 0, ‘I’(z) < 0. In general, for an attractive fluid near an at- tractive wall an increase in density is exhibited near the wall depending on the magnitude of ‘I’(z). If the wall—fluid forces are stronger than the fluid-fluid forces, the fluid Will wet the wall, otherwise there will be a rarefaction of gases near the wall. To represent the in- teractions of the adsorbate and adsorbent we have chosen the partially integrated 10-4 potential model [Lee, 1988] . 4 1 ‘I’(z) = 47rpamef,o,3[5:—:f— m— -Zx—4] [2-55] 1’1 I Where pm," = 0 382 atoms/ A2 , x,- is the intermolecular distance between fluid molecular Centers and the ith plane of solid molecules, where we have truncated the interactions at the fourth plane of solid atoms, where the interplanar spacing is 3.35 A. sf, is the well depth of the fluid-solid potential, and is obtained from the Lorentz—Berthelot mixing rules Which state that efs = (8,; 8:3)0'5 and of, = 0,5 (of, + 0'”). Table 2-1 summarizes other pa- r atneters used in this work. Note that Efl‘ is not tabulated because the fluid-fluid contribu- tion is calculated from Equations 2-46 - 2-48 and not directly from 8,]. Steele [1974], 47 Table 2-1: Molecular properties used in SLD model System 6/] (nm) of, (nm) £,..//< (K) Ethylene-Graphon 0.422 0.38l 450. 23 Krypton-Graphon 0.35 0.345 225 48 Nicholson and Parsonage [1982], and Lee [1988] offer a further discussion of fluid-solid potentials. Figure 2-10 shows adsorption isotherms that were calculated using the 104 po- tential of Equation 2-47 and the parameters for ethylene and graphon in Table 2-1. The curves are qualitatively similar to Figure 2-2. A knee is present, the magnitude of which can be increased by increasing the solid-fluid interaction energy 8,, (Figure 2-10 also shows calculated adsorption isotherms for condensed phases at subcritical temperatures above the bulk vapor pressure which are not present in Figure 2-2). Some characteristic features of adsorption of near-critical fluids are the cusp-Shaped isotherms seen near the critical point, and the 'cross-over' of adsorption isotherms at different temperatures. Con- sider the effect of temperature at a fixed pressure below the critical pressure. Below the critical pressure, adsorption decreases with increasing temperature, but this trend is re- versed above the crossover region at pressures above the critical pressure. The crossover of the isotherms is eliminated by plotting the calculated surface excess with respect to cal- culated bulk density as shown in Figure 2-11. The reason for the cusp-like behavior of a supercritical isotherm in Figure 2-10 can be understood from the calculated density pro- files shown in Figure 2-12. Below the critical pressure region, the adsorbed layer in- creases in thickness faster than the bulk density increases. Above the critical pressure re- gion, the bulk fluid becomes increasingly incompressible, and the bulk fluid density ap- Proaches the density of the adsorbed fluid, causing the surface excess to decrease. 49 30 25 -- Excess (ll mol/mz) 61' B —l O I . ---263.15K '1 : ------ 273.15K , .' —283.15K ,' ----293.15K I ----303.15K 20 40 60 80 100 Pressure (bar) Figure 2-10: Adsorption of Ethylene on Graphon 120 Excess (ll mollmz) 25 50 20 -~ —I 01 l d O l -—-283.15 K --- 293.15 K ------ 303.15 K ..... 'O l l l l 0 l I I I I I ‘ 0.002 0.004 0.006 0.008 0.01 Density (moles/ems) Figure 2-11: Adsorption of Ethylene on Graphon 0.012 Density (moles/ems) 51 0.02 - - - 38.83—bar 0.02 -- ------ 42.76 bar —47.26 bar 00] __ ----so.04 bar ' - - - — 54.22 bar 0.01 -- 0.01 -— ‘5: 0.01 -- \ ................... _ l ‘, I : 0.01 —- I g 1 '. a. I 2‘ ----------- [1m —-— \\ . \\ ________________________________________ 0.00 —— """""""""" 0.00 I : I 0 5 10 15 Zle Figure 2-12: Density Profile of Ethylene on Graphon at 283.15 K 20 52 Type III isotherms are normally observed when the attraction between the solid and fluid is weak. Figure 2-13 demonstrates the capability of the SLD model for predic- tion of Type III isotherms. In this case, the fluid parameters are for ethylene, but the magnitude of the 8,, has been decreased. The smaller 8,, results in elimination of the knee and yields a Type III adsorption isotherm. At very low pressures and temperatures, a dis- continuity in the adsorption isotherms is predicted (not shown here), indicating a phase transition in the adsorbed phase similar to those discussed for a two-dimensional model [Ross and Olivier, 1964]. At high pressures, when the bulk fluid exists as a liquid, the SLD model can predict a negative surface excess when Efs is small. Negative surface ex- cesses have also been predicted by Sullivan [1979]. Figure 2-14 shows some predicted adsorption isotherms of krypton on graphon, at temperatures far above the critical tem- perature. When compared with experimental data shown in Figure 2-3, the predictions are again qualitatively correct. This model predicts some gas-liquid transitions on the surface, similar to the transitions seen in the experiments of Thomy [Bienfait, 1980]. 53 35 ——203K ---233K ------ 263K 25 ~- Excess (ll mollmz) 51' B ._.I O l T “’— -—-‘-"— ...-.....................C.........-.. Pressure (bar) Figure 2- 13: Weak Adsorption of Ethylene on Graphon 54 Figure 2-14: Adsorption of Krypton on Graphon 14 12 i 10 ‘— II” \\\\ “A [I ........................ \. g z’/ \K‘;-- 0 8 -- x’ —— E I, ................ \ =- I .o ----- V / o .d’. m I .'. "/ m 6 "' /.’ ’z" a) .- .- 0 I. .1 . x 0‘ I In .' ,o’ 4 '1 [O i ——253K J ---273K 2 ------ 298K —348K - --373K 0 i I I 4 4 I I 0 20 40 60 80 100 120 Pressure (bar) 55 SLD Model - Discussion The SLD model builds on the approaches of the simple theories such as Langmuir, BET, two-dimensional equations-of—state and the van der Waals approach of Barrer and Robins. By treating the fluid with a van der Waals equation with a suitably modified a, the SLD model allows for interactions between adsorbed molecules at various distances from the wall. All of the models mentioned above (including the SLD) assume an energetically homogeneous surface. The effect of heterogeneity will be pronounced at extremely low pressures and coverages, where the high energy sites are unfilled. To represent a hetero- geneous surface, one could fit the potential to adsorption in the Henry’s law region, and then work with a pseudohomogeneous surface. The limitations of the model can be attributed to: (i) the lack of structure in the fluid; (ii) the use of the local density approximation; and (iii) the use of the van der Waals equation to describe the fluid properties. Since we are using the van der Waals equation and a mean field approach, we do not predict any of the fluid structure seen using either computer simulations or density functional theory [Snook and Henderson, 1978; Vander- lick et al., 1988; Kierhk and Rosinberg, 1991]. This model cannot describe discrete fluid structure or be used to study packing near a wall. The use of the local density approxima- tion also results in the physically unrealistic prediction of abrupt vapor-liquid interfaces. Despite the fundamental theoretical limitations of the model, since model predictions mimic experimental trends, this model may be acceptable for engineering calculations. In order to estimate the errors introduced by the use of the local density approximation, Equation 2-42 was solved keeping the density inside the integral, leading to an integral 56 equation (IE) [Pyada, 1994]. When compared, solutions to the SLD model and IE ap- proach showed differences which were dependent on pressure, temperature and magnitude of the gas-solid interaction potential. For the adsorption of ethylene on graphon, the SLD model tends to underpredict adsorption (relative to the integral equation (IE)) by about 10-20% between 0.9 < Tr < 1.1, except where the surface excess increases steeply at high pressures. At pressures near 1 bar, the differences are less than 1%. These comparisons show that the local density assumption can provide a reasonably approximate solution to the adsorption problem. A limitation of the van der Waals-based SLD model is the fact that predicted magnitudes of adsorption and vapor pressures are incorrect. For accurate prediction, the equation-of-state must be able to correctly predict the vapor pressure and liquid density. The inaccuracies of the van der Waals equation in this regard are well documented. In order to represent the fluid properties any equation-of-state can be used, provided the re- pulsive and configurational contributions can be separated for use in Equation 2-39. The selection of the van der Waals equation as the basis of this work was due to the fact that it is the simplest and most easily adapted equation-of-state with a theoretical basis. The ob- jective of this work is to demonstrate that the proposed approach provides qualitative predictions with the van der Waals equation, and lay down the framework required to use a more accurate equation-of-state. A comparison of Figures 2-2 and 2-10 shows that the SLD model exhibits poor prediction of the vapor pressure for subcritical isotherms, but - does better at predicting the pressure of the maxima in the supercritical isotherms. This is because the van der Waals parameters a and b were obtained from the critical temperature 57 and pressure. The 8/, in Table 2-1 have been selected to provide semiquantitative fit to the magnitude of experimental adsorption. Cubic equations are widely used in industry, and a method that adapts them to the adsorption problem could find widespread use in process calculations. In the next chapter we will see how changing the fluid equation-of-state significantly improves the quantita- tive modeling of adsorption. We will also show the extension of this model to predict Type I, IV and V isotherms, and multicomponent mixtures. CHAPTER 3: QUANTITATIVE MODELING OF ADSORP- TION AND EXTENSION OF THE SLD APPROACH TO PREDICT CLUSTERING INTRODUCTION In the previous chapter, we presented an engineering model that adapts the van der Waals equation-of-state to describe the fascinating behavior seen in the experiments of Findenegg [1983]. The fluid-solid potential was superimposed on the van der Waals equation-of-state, and the configurational energy integral in the inhomogeneous fluid phase is simplified with a local density approximation. While this model does a good job of qualitatively describing the adsorption behavior, quantitatively it does not predict the isotherms very well. The primary reason for this shortcoming was that the van der Waals equation does not predict accurately the vapor pressure and density of the adsorbing fluid. In this paper we use the Peng-Robinson equation to describe the fluid properties, which significantly improves the model predictions relative to the experimental results. An adsorption model may also be adapted to describe clustering in supercritical fluids [Lee et. a1, 1991]. Clustering is a phenomena that occurs in supercritical fluids whereby a large number of solvent molecules collapse around each solute molecule [Debenedetti, 1987; Eckert et. a1, 1986], and the effect is most pronounced near the criti- cal point. One of the first experimental studies on clustering involved the measurement of partial molal volumes near the critical point for solvents such as carbon dioxide and ethyl- ene around solutes such as naphthalene [Eckert et. al., 1986], demonstrating that the par- tial molal volumes become extremely negative near the critical point. Debenedetti [1987], 58 59 using a fluctuation analysis converted these negative partial molal volumes to number of solvent molecules that cluster around the solute. Kim and Johnston [1987 a, b] state that at 300.5 K and 79.8 bar, the partial molal volume of naphthalene in supercritical carbon dioxide at infinite dilution is -7800 cc/mol, which corresponds to the condensation of 80 solvent molecules around a solute molecule. In a related study, Johnston et. al. [1987] show that there is a shift in the solvatochromic data of phenol blue in ethylene indicating an increase in the number of solvent molecules around a solute. Wu et. al. [1990] provide integral equation calculations on other supercritical systems that also suggest that cluster- ing does occur in supercritical mixtures. Brennecke and Eckert [1989] measured the intensity ratio of the fluorescence spectra to probe the local density of dilute organics in pure supercritical fluids. The in- tensity ratio for naphthalene in CO; at 4 K above the critical point was much greater than at 14 K above the critical point, indicating the clustering effect near the critical point. Such an increase was also demonstrated for systems such as pyrene-CO; and pyrene- ethylene. Lee et. a1. [1991] have reviewed both the integral equation and spectroscopic measurements that give evidence of clustering in supercritical fluids. Clustering is used to explain the unusual behavior in supercritical fluids such as increased solubilities, synergistic effects of mixed solutes and entrainer effects. The forces acting on the fluid molecule which result in this behavior are the fluid-fluid and fluid-solute intermolecular forces. On a molecular level, there is a similarity to the fluid-fluid and fluid-adsorbent intermolecular forces in physical adsorption on solid surfaces. In cluster- ing, this leads to an increase in density of the fluid molecules in the region around the sol- 60 ute similar to the increase in density around the adsorbent in physical adsorption, and the fluid can be modeled in the inhomogeneous system. In this chapter, the results of this simplified local density (SLD) model using the Peng-Robinson equation-of-state are given. The basic concepts of the SLD approach have been discussed in the previous chapter. The modifications made to this model to in- corporate the Peng-Robinson equation, and model predictions of experimental results for adsorption on flat walls are discussed in detail in this chapter. The modifications made to this model to predict the clustering phenomenon are also discussed in detail. The model is capable of predicting the density profile and number of solvent molecules around a solute to indicate long-ranged interactions, giving evidence of clustering around a solute near the critical point. The model is shown to correlate with fluorescence spectroscopy measure- ments. MODEL DEVELOPMENT The approach of the model is the adaptation of a cubic equation-of-state to define the properties of a fluid that is in the external potential field of an adsorbent. First, we de- velop the model for flat solid adsorbents (incorporating the Peng-Robinson equation) and then generalize the results for clustering. A fluid molecule at any position from the solid interacts with both the solid and fluid molecules. Therefore, the chemical potential of the fluid may be written as a sum of the fluid-fluid and fluid-solid interactions Il(z)=llb =ug(2)+u,.(2) [3-11 61 where ttfflz) and 11st z) are the chemical potentials due to the fluid-fluid and fluid-solid interactions, and z is the distance normal to the surface of the solid, and the subscript b stands for bulk properties. If the fluid-solid potential is given by 111(2) on a molecular ba- sis, then on a molar basis Hf, (Z) = N A‘I’(Z) [3'2] For a non-ideal bulk fluid, the chemical potential jib of Equation 3-1 is related to the fugacity as u. = #0 + R T ln(-£42] [3-3] 0 Details of the development of this model are given in the previous chapter. Using the Peng-Robinson equation-of-state, can write an expression for the bulk fllgacity in terms of molar volume. ln(-1%):(Zb - 1) - ln(Zb - B) - A, ln(z, +2414 B] {3.4} 2.828 B z, - 0.414 B where 62 P . . .. Z, = RV; IS the compressrblllty factor a P Ah R2bT2 3:21: RT P= RT ab vb -b_ vb2 +2bvb -b2 where a, b are the constants of the Peng-Robinson equation [1976]. Rearranging Equations 3-2 - 3-4, the component of the fluid fugacity due to fluid- fluid interaction is ‘1’( ) f 17 (z) = f), “PL—[TL] [3-5] Using the PEng-Robinson equation, the fluid-fluid fugacity at any point becomes _ b _ 0(2) V(z) lnlffld] — v(z)—b v(z)2 +2bv(z)—b2 RT - 1,1[M] _ 0(2) 1nl:v(z)+2.414 b] 13-6] RT 2.828bRT v(z) - 0.414 b 63 where v( z) is the molar volume at any point z. a(z) is calculated using the SLD approach, as given in the previous chapter, assuming a fluid-fluid pair potential oc f6 and is given by a(z) = a, —5— + ii for 0.5 s —z— s 1.5 [3-7] 16 166, of 1 z a(z) = ab 1- for 1.5 S — S 00 [3-8] 8(—Z— — 0.5)3 0 x G where of is the diameter of the fluid molecule. In modifying this model to arrive at Equations 3-6 - 3-8 for the Peng-Robinson equation assumptions are necessary to evaluate the configurational integral. Returning to the integral from which the equation-of state a parameter is based, the attractive chemical potential is given by u... = jurfifldv 13-91 V Where V denoted the volume of integration, (p(r) is the two-body interaction potential, and 8( r) is the radial distribution function. 64 In the SLD model, the fluid at a point z is treated as a homogeneous fluid at a density of p(z). Solving Equation 3-9, the attractive pressure term of the van der Waals equation for a homogeneous fluid becomes -a/v2'. In the case of the Peng-Robinson equa- a v2+2bv—b2° tion, the attractive pressure term for a homogeneous fluid is — The Peng- Robinson equation, as an empirical equation-of-state, has an unknown equivalent form of the integral given by Equation 3-9, and therefore some assumptions are necessary to ap- proximate this integral. The term v2 in the denominator of the attractive part of the van der Waals equation is a result of the volume derivative of Equation 3-9 . The term v2 + 2bv- b2 in the Peng-Robinson equation may be treated as an empirical equivalent of the term v2 of the van der Waals equation, and can be subject to an analogous SLD approxi- mation. If the density is uniform up to the surface of the adsorbent, then we know that at this limit 11qu = rib/2, resulting in a..." = al./2 for the van der Waals equation. The Peng- Robinson equation should have the same limits. Also, the functionality of the a/ab should have the same qualitative curvature in both the van der Waals and Peng-Robinson equa- tions. Since the true form of Equation 3-9 is not known for the Peng-Robinson equation, as a first approximation, we use the van der Waals ratio of a(z):ab for the Pong-Robinson SLD model, and retain v2 + 2bv- b2 as the denominator of the attractive pressure term, Which results in Equations 3-6 - 3-8. The other pair potential needed for this model is the fluid-solid potential. The ad- Sol‘bate-adsorbent interactions are given by the integrated 10-4 Lennard-J ones potential ' [Lee, 1988] as 65 0‘}. 1 4 1 ‘I’(z)=47rpa,m8fi0'fi EFT-2'27 [3-10] 1 i=1 i For modeling clustering in supercritical fluids, the approach is essentially similar, with the exception of the configurational integral calculation. The wall in adsorption onto a flat surface is replaced by a solute molecule with the solvent molecules clustering around it. Geometrically, this means replacing a planar surface with a spherical particle. The co- ordinate system is changed from rectangular geometry to spherical geometry and the fu- gacity, molar volume and the Peng-Robinson a term become functions of r, the radial dis- tance from the center of the solute molecule. The corresponding a( r) become '8 2 , 2 , 1 ‘ 3+;(r-r “30’ ( + ’)3 r r —a(r) =3 ’ for 0.50, s r-91 s 150, [3-11] +— — _ r {7’2 +r2 —277‘” r’2 +r2 +2rr’}_ P16 26) 1 203, 1 —- + a(r)_i 3 3 (r—r’f 3 (r+r')’ ab 16 +03{ 1 l } r r’2+r2—2rr’ r’2+r2+2rr’_ for 1.50, s r--62—‘ s cop-12] Where _(o,+o,) I 66 r’2 -l-r2 —O'2 n_ f r — I 2r where the subscript f stands for fluid, and the subscripts for the solute. Details are given in the appendix 1. The solute-solvent interactions are given by the Lennard-Jones potential 0.12 0.6 ‘I’(r) =4e,,(-;’2i--r7" [3-13] The component of the fluid-fluid fugacity due to fluid-fluid interactions, i.e. the equivalent form of Equation 3-5 for clustering is ‘i’( ) fflm = fbexp[- k; ] [344] Using the Peng-Robinson equation-of-state, the fluid-fluid fugacity can be expressed Similar to Equation 3-6, except that the molar volume, a and fugacity are functions of r instead of 2. Algorithm for solving density profile To calculate the surface excess or cluster size, the density profile must be known. The density can be evaluated by first calculating the bulk fugacity and density from a given 67 temperature and pressure. Knowing the fluid-solid potential, the fugacity at any point can ' be calculated by Equation 3-5 (physical adsorption) or 3-14 (clustering). The density is then calculated iteratively using Equation 6 or by the modified form using v( r) and a( r) for clustering. Once the density profile is known the surface excess can be calculated by equa- tion 2-1. Similarly, for clustering, the cluster size (Nu) can be calculated as N" = f(p(r)— pb)dV [3-15] The FORTRAN programs for calculating the adsorption isotherms and cluster sizes are available in appendix 2. RESULTS The diameters of the fluid, solid and solute are the Lennard-Jones parameters as tabulated in Reid et. al [1987] and in Lee et. al. [1991]. The Of; is calculated as the arith- metic mean of the fluid and solid (or solute in clustering), while the fluid-solid potential (83) is used as an adjustable parameter to adjust the magnitude of adsorption. The values Used for all the parameters are summarized in Table 2-1 or 3-1. Figure 3-1 shows the surface excess of ethylene adsorbing onto graphon at various Sub— and supercritical temperatures, and over a wide pressure range. This model is capa- ble of quantitative fits for the adsorption isotherms over this entire pressure and tempera- ture range, using a single parameter. The model also does a good job of accurately pre- dicting the pressure at which the peaks occur, showing the characteristic cusp-like behav- Table 3-1: Molecular properties used in the SLD model 68 System odnm) 0, (nm) Cg (nm) EfS /k (K) Ethylene - Graphon .422 .34 .381 140 Krypton - Graphon .3655 .34 .35275 95 Propane - Graphon .51 18 .34 .4259 110 Argon - Graphon .3542 .34 .3471 68 Methane - Graphon .3758 .34 .3579 107 Carbon dioxide - Naphthalene .3794 .6199 .49965 800 Carbon dioxide - Pyrene .3794 .714 .5467 850 Ethylene - Pyrene .422 .714 .568 800 Excess (It mollmz) 69 Pressure (bar) Figure 3-1: Adsorption of Ethylene on Graphon I20 45 1:1 1:) 263.15 K 40 __ A A 273.15 K A o 283.15 K 0 293.15 K 35 “r A g - 313.15 K o A o 30 " A ° 1:1 A o l} o 25 "" I I o; I 20 -- 9! °.-" 4’ 8" 03968 15 - I f" 9”"\ 8 03’ o ‘ O ’6’ *‘g —!-‘~ 10 - 1,..-"""o \x “x- ’ 90 “ ‘\..~ “‘~-. °o o -~a::”f~-—_ 5 ‘ .7 ° 0 o O” i i 0 u I I I I I (I 20 40 60 80 HI) 70 ior seen near the fluid critical point. The model predicts the correct temperature- dependent crossover of adsorption isotherms in the supercritical region. One of the pri- mary reasons that the Peng-Robinson SLD model does a better job than the van der Waals SLD model is that the Peng-Robinson equation predicts more accurate density values near the solid surface, and a more compressible adsorbed layer. A comparison of experimental saturation densities for ethylene [Angus et. a1, 1974], and both these equations [Pyada, 1994] shows that while the van der Waals equation underpredicts the saturation densities by 80-90% between 211 and 260 K, the Peng-Robinson equation is within about 6%. van Megen and Snook [1982] have done Monte-Carlo simulations of Lennard- Jones mole- cules (that simulate ethylene) on graphon at supercritical temperatures [8] (T IT c = 1.03) (see Figure 2-5). Their results significantly overpredict the experimental values of Fin- denegg, and their maxima is around 42 bar, as opposed to the experimental value of 58.7 bar, while the SLD model has a maxima at 57.8 bar. Figures 3-2 and 3-3 Show the adsorption isotherms of krypton onto graphon at supercritical temperatures, and propane at subcritical temperatures. The krypton iso- therms 253 and 273 K crossover at around 140 bar. Once again, the model predicts this crossover at the correct pressure. This model was also tested on gases like argon and methane (Figures 3-4 and 3-5). While, the model predictions for argon were good, the model did not do a very good job of predicting the isotherms for methane, the reason maybe that the Peng-Robinson equation does not do a good job of predicting the fluid properties of liquid methane [Prausnitz, 1980]. The PVT behavior of these different gaSes and the error in predicting density values is shown in appendix 3. H 14 a D :1 1:1 12 "' n D a n o o 0 o 10 -- ° ° - 01‘ o ’, """"""" ~-‘o E n o ’z” a = x . . O 8 -- l/’ A ........ 0 g. D °/ A .f L--—""‘" __-,_.....-._. V II 9 ,I”’ " .— "- .-.‘_.._ _. m 9’ .. . ' ". ‘ ‘— 8 6 ‘T l I 4' .,( _,r '3 ‘ A O I l.’. ,': ° A ‘ ‘ I '0 ./ .o ‘ 4 a- 79' / '.’ ‘ ‘ 13253.15K ..' . , "Ii/’3. o ‘ 0 273.15 K 'a'/k . ‘ g A 298.15 K 2 -:'.' ‘ l323.15K :5"; 0348.15K t 1373.15K 0 I I I 0 40 80 120 160 Pressure (bar) Figure 3-2: Adsorption of Krypton on Graphon 72 30 i . I . I I 25 -- I i I I l "O l . l 18 "A 20 -— '1 I; g ’ a — 1:1 I 8. O I I E i ,0 .3- 15 I’ o" v) I 00/ 8 /' O 99/ X W' Lu 10 a 'wo’ A“! ,0"? A “0’99" 1 13273.15 K { 0323.15 K 41343.15 K 0363.15 K 20 30 Pressure (bar) Figure 3-3: Adsorption of Propane on Graphon 4O Excess (u mol/mz) 73 9 8 -- 7 .- 6 -— 5 .. 4 -- 3 -- I .' Fl . "7' 0.; n253.15K .. ' O 2 . /'/ «1273.15 K 1",” 4298.15K 0 323.15 K O -- I I I 0 40 80 120 Pressure (bar) Figure 3-4: Adsorption of Argon on Graphon I4 74 12~ IO- Excess (It mollmz) a 253.15 K o 273.15 K A 298.15 K 0 323.15 K l I l I l l 80 l 20 Pressure (bar) Figure 3-5 : Adsorption of Methane on Graphon 160 75 Figures 3-6 - 3-12 describe the clustering phenomenon. The density profile and the excess number of solvent molecules within a sphere of radius R around a solute are plotted in Figure 3-6. As shown in the figure, the density goes through a sharp increase near the surface of the solute but quickly drops to the bulk value. The excess number of solvent molecules is slightly negative very near the surface before increasing sharply. Such a trend has been shown in the integral equation calculations of xenon in neon [Wu et. al, 1990] (Figure 3-7). The figure also shows that the excess number of solvent molecules increases sharply up to 5 molecular diameters, and then slowly till 9 or 10 molecular di- ameters, showing the long-ranged effect of this clustering phenomenon. This trend has also been reported by Wu et. al [1990]. Figures 3-8 & 3-9 shows the results for clustering of C02 molecules around naph- thalene at 2 K, 4 K and 14 K above its critical point. As seen in Figure 3-8, the cluster size increases sharply near the critical point, but drops quickly as we move away from the critical point. The results of the fluctuation calculation of Debenedetti [1987] on the same system at 308 K are also plotted in Figure 3-8. The model predicts the correct density at which maximum number of C02 molecules cluster around naphthalene, and does a fair job of predicting the number of CO; molecules. The model also predicts a slightly broader density range over which this clustering phenomenon occurs. However, the model does accurately predict the increase in the number of C02 molecules around a single naphtha- lene molecule as the critical point is approached, and also the variation of this peak with temperature and pressure. When this cluster size is plotted against bulk density, cross- Nex 76 80 30000 L-25000 l L r200m L ~15000 «10000 Nex ' - " - P -10 4. I I 0 0.00 5.00 10.00 15.00 20.00 Rlo Figure 36: Density and Excess number of solvent molecules within a Sphere of Radius R around a Solute (sumowfi) d 77 a... ....... 0' .o o o .o 701 10‘ I [g,2(r) - l] ‘Irtr2 dr ((511)3 0 L l O I H O 13 Is L10“ C U11 F1 QUre 3-7: Number of Excess Solvent Molecules Within a Sphere of Radius L Around a Solute A: Xenon in Neon (Attractive) F1: Neon in Xeonon (Repulsive) P: Pure Lennard-Jones 78 100 D - - - 306.2 K 303.33 K ll a 308.38 K,inferred from expt. ll H I: ------ 318.2K I ul- - 1- 80 100 120 140 Pressure (bar) - Figure 3-8: Clustering of CO2 around Naphthalene 160 79 .' - - - 318.2 K 90 __ ------ 306.2 K -' ‘ a 308.38 K, inferred from expt. 308.38 K Nex 8 0 : l l n 4000 8000 12000 16000 20000 Density (gmollm3) Figure 3-9: Clustering of C02 around Naphthalene Localized Cluster Size 80 0.38 58 -- 4:- 0.36 54 -— .. 034 w -- A .. 0.32 A A AA 46 " -—303.2 K ._ 03 ------ 313.2 K a Data - 308.2 K A Data - 313.2 K 42 i i i 023 4000 8000 12000 160m 20000 Density (gmollma) Figure 3-10: Comparison of Cluster Size with Fluorescence Spectroscopy - CO2 around Naphthalene VII I-I Localized Cluster Size 88 81 1.1 32 j- a .T '| 76 T -- 0.9 70 -- 64 T “ 0.8 58 uh -- 0.7 52 -- -.—303.2 K ------ 323.2 K L' 0-6 46 " a Data 303.2 K . A Data 323.2 K 40 J : : : 0.5 0 5000 100(1) 15000 20000 Density (gmollm3) Figure 3-1 1: Comparison of Cluster Size with Fluorescence Spectroscopy - CO, around Pyrene Elll-l Localized Cluster Size 82 72 0.8 68 -- -- 0.75 64 .... 60 -- 0.7 56 -- . " 0-65 D ,5. ' 52 -- ° a -- 0.3 43 -- ° a —234.2 K ------ 303.2 K “ 0-55 44 "" a Data - 234.2 K a Data - 303.2 K 40 1 ' i 1 i 0.5 0 3000 6000 9000 1 2000 1 5&1) Density (gmollma) Figure 3-12: Comparison of Cluster Size with Fluorescence Spectroscopy - Ethylene around Pyrene Ell l-l 83 overs of Figure 3-8 disappear, and the maxima appear at the same bulk density (Figure 3- 9). The fluorescence intensity ratios are an indication of the total number of molecules around the solute within a short range, and hence the total amount within the first two . o. o, . molecular diameters [—'— S r S -—‘ + 20 f] outsrde the solute was chosen as a measure of 2 the local density. Figures 3-10 - 3-12 show the total number of molecules around a solute within the first two molecular diameters against bulk density. This is cross-plotted against the fluorescence signal of Brennecke [1989], Brennecke and Eckert [1989] and Brennecke et. al. [1990]. The increase in the intensity of the signal near the critical point suggests that the solvent molecules cluster around the solute. For naphthalene, the stable signal is the ratio between fluorescence peaks 1 and 4, while for pyrene it is 1 and 3. These figures show the predicted total number of C02 molecules clustering around naphthalene and py- rene, and ethylene clustering around pyrene. The model predicts the trends seen in the fluorescence spectra. For CD; clustering around pyrene and naphthalene at 308.2 K, the data show a sharp increase at low densities, then becomes rather flat before increasing again at high densities. Also, at higher temperatures, the flat region is rather small. Once again this trend is captured by the model. This difference is the trend of the fluorescence spectra is even more pronounced for ethylene clustering around pyrene. The figures also show that the intensity of the ratio of the fluorescence spectra and the excess number of C02 molecules around pyrene is more than around naphthalene. The fluid-solvent inter— action parameter is higher for pyrene than for naphthalene, and pyrene being a larger molecule provides a larger surface area for the solvent molecules to collapse upon. Com- 84 paring the C02-pyrene and ethylene-pyrene systems, pyrene attracts a larger number of C02 molecules than ethylene. The reason may be because of the bigger size of the ethyl- ene molecule which sterically limits the number of ethylene molecules that can surround the solute. In all the cases presented here, the model predicts greater clustering near the critical point which is in agreement with the spectroscopic data. DISCUSSION The SLD model uses the fluid-fluid and fluid-solid interaction parameters to model adsorption over wide pressure and temperature ranges. It characterizes the fluid mole- cules using cubic equations-of-state. Predictions of the SLD model using the van der Waals model were shown earlier [Rangarajan et. al, 1995, chapter 2]. Here, we use the Peng-Robinson equation to describe the fluid, thus significantly improving the predictions of the fluid properties. By comparing Figures 3-1 - 3-2 with the results seen in chapter 2, the significant improvement is obvious. However, for any particular system, the equation- of-state that does the best job for representing its properties may be chosen, and the SLD approach adapted for that purpose. The fluid-wall potential is used as an adjustable parameter in this approach. It is fitted to any one of the different temperatures given in the data, and then the same poten- tial is retained for the other temperatures. This makes the model a semi—predictive model, in that, while one of its parameters needs to be fitted to experimental data at a particular temperature, it then becomes a predictive model to describe the adsorption characteristics at other temperatures. The geometric mean of the Lennard-J ones potentials tends to un- 85 derpredict the values of the surface excess. We have consistently used a fluid-solid inter- action parameter that is about 1.8 to 2.2 times the geometric mean. The SLD model superimposes the Lennard-Jones potential for solute-solvent in- teractions on the Peng-Robinson equation of state that describes the properties of the su- percritical solvent. If the solute is attractive, then the density of solvent molecules near the solute increases resulting in the phenomenon of clustering. This is very similar to the adsorption of a fluid on to a flat wall, except that the wall-fluid interactions are slightly different than the solute-solvent interactions. As shown by the results of both these two cases, there are many similar behaviors between the two cases. The crossover of the ad- sorption isotherms is also seen in the case of clustering (Figure 3-8). This crossover dis- appears when the excess is plotted against bulk density (Figure 3-9). The peak in the su- percritical adsorption isotherms decreases as we move to higher reduced temperatures. This phenomenon is also seen in clustering, where the number of excess solvent molecules drops sharply as the temperature is increased from 2 to 4 K above the critical point. One of the differences is that below the critical pressure, the surface excess of a fluid adsorbing on a flat wall increases gradually and then drops sharply as seen at 283 and 293 K for eth- ylene (Figure3-l). However, in clustering, as the fluctuation calculation of partial molal Volume data indicates, the number of molecules increases sharply as the critical pressure is approached and decreases rapidly past the critical pressure. This model’s predictions for Clustering are more in line with the adsorption data on a flat wall i.e., the excess number of Solvent molecules increases more gradually below the critical pressure, and then drops Sharply after the critical pressure. This may be an artifact of the SLD assumptions. [it he H0 D0 T10 [11. if: fen 10] ill 86 Wu et. al. [1990] showed that for the attractive mixture of neon clustering around xenon, there is a negative N“ close to the surface of the xenon molecule (Figure 3-7). The SLD model also predicts such a negative value close to the surface. At the surface, the attractive and repulsive parts of the Lennard-J ones potential cancel, and the solvent molecules are not attracted to the solute, and this negative value is the result of the sol- vent-solvent interactions. By choosing the correct solute-solvent potential parameter, this model may be used to describe a repulsive mixture where there is depletion of solvent molecules e. g. xenon molecules cavitating around neon. The SLD model serves as a good first approximation of the adsorption problem. The use of the Peng-Robinson equation helps resolve one of the limitations of the previous SLD model viz. use of an equation that can better represent the fluid properties, thus sig- nificantly improving the capabilities of this model. High-pressure adsorption and clustering in supercritical fluids are phenomena that are of importance in various industries. A simple engineering model to describe these dif- ferent phenomena will enhance their applications in industry. Modifications of this model to predict adsorption in slits and pores, and for describing adsorption of mixtures onto different surfaces are discussed in the next chapter. {1 CHAPTER 4 ADSORPTION'OF PURE GASES IN SLITS AND PORES, AND ADSORPTION OF BINARY MIXTURES INTRODUCTION in two previous chapters, we presented a new model, called the Simplified Local Density (SLD) model, for describing physical adsorption of gases on flat walls using cubic equations-of-state. In this model, the fluid-solid potential is superimposed on a cubic equation-of-state, and the configurational energy in the inhomogeneous fluid phase is simplified with a local density approximation. In the first chapter, we used the van der Waals equation, which gave qualitative predictions of the experimental data of Findenegg [1984]. The van der Waals equation was used because it is the simplest fundamentally sound equation-of-state and provided a basis for qualitative studies. This model predicts Type II and III isotherms for adsorption on flat walls, shows the characteristic cusp-like behavior and crossovers seen near the critical point. In the second chapter, we modified the model to incorporate the Peng-Robinson equation-of-state, which represents the fluid properties more accurately, including vapor pressure and compressibility, and showed dramatic improvement in the quantitative representation of experimental measurements of Ffindenegg. Most commercial adsorbents are porous, activated carbon being one of them. Ad- sorption in pores and the phenomenon of pore filling among carbon based adsorbentshas received considerable attention [Dubinin, 1975]. Gregg and Sing [1982] have classified pores into three categories, macropores (50 - 100 nm), mesopores (2 - 50 nm) and micro- 87 88 pores (<2 nm). Adsorption properties of microporous solids enjoy wide spectrum of ap- plications including chromatography and other separation processes, filtration, industrial effluent cleanup, ion exchange, biological applications etc. [Gubbins, 1990; Pendleton and Zeetllemoyer, 1984]. In microporous solids, adsorbed fluids show many unusual proper- ties such as preferential adsorption of certain fluid components, chemisorption at particu- lar sites, hysterisis effects and a variety of new and unusual phase transitions [Gubbins, 1 990]. Dubinin [1975] says that all existing theories of physical adsorption proceed with the same physical image in describing adsorption in porous and nonporous adsorbents. Design of columns for thermal swing adsorption and pressure swing adsorption require the simultaneous solution of partial differential equations for the material, energy and momentum balances describing the dynamics of adsorption in columns, in conjunction with the kinetic and equilibrium pr0perties of the adsorbents [Sircar and Myers, 1985]. The models for multicomponent adsorption isotherms are discussed in chapter 2. There are plenty of experimental and molecular simulation data available for adsorption of pure gases and mixtures onto slits and pores at low pressures [van Megen and Snook, 1982, l 985; Barton et al., 1984; Powles et al., 1988; Valenzuela and Myers, 1989; Tan and Gubbins, 1990; Kieer and Rosinberg, 1992]. In this paper, we extend the SLD model to describe adsorption of pure gases and binary mixtures in confined spaces, i.e. slits - two infinite parallel flat surfaces and cylindrical pores. 89 MODEL DEVELOPMENT - PURE GAS Slit-like Pores These can be thought of as two parallel semi-infmite plates separated by a small distance (Figure 4-l). Any particle in between the walls will be subjected to attractive forces from the two walls. The distance between the two walls may be quite small (<2 nm), which implies that only particles smaller than this will be adsorbed i.e. exclusion will be important. The chemical potential of the fluid in the vicinity of a wall may be written as a sum of the fluid-fluid and fluid-solid interactions, as described in the previous two chapters. Recapping from the last chapter, using the Peng-Robinson equation-of-state, we can write an expression for the bulk fugacity in terms of molar volume [Sandler, 1989]. The fluid-fluid fugacity at any point is an analogous expression b _ a(z) V(z) _ v(z) — b v(z)2 + 2bv(z) - 1)2 RT ln[v(z)-b] _ a(z) ln[v(z)+2.4l4 b] RT 2.828bRT v(z)-0.414 b 1“”.ng = [4'1] where v(z) is the molar volume at any point z, a and b are the constants of the Peng- Robinson equation [Sandler, 1989; see previous chapter also]. a(z) is calculated using the SLD approach after suitable modifications, and details are given in appendix 1. The equa- tions for a(z) in slits are 90 plane of symmetry 9(2) p: Potential Energy Slit-Like Pores 1 Figure 4- 'he ten 91 “(0:2 _Z_+.5__ 1 3 , for assists [4-2] ab 8 017 6 L-Z 0’7 3 ———0.5 a(Z)_§ §_ 1 _ l a -8 3 3 _ 3 , b 3[__Z__05] {L 2—05] or of a [4-31 for 153-3—3—5—15 o, 617 0(2):; L-z+%_ 1 3 , for BL_1.5_<_O_Z_gal;.—0,5 [4-4] a 0' ” 1’ 3[—5——0.5J ” f H 0'17 - Wher e L+o, is the distance between the centers of the atoms of the two surfaces. Note that L has to be at least 3 Gfl' in length for the above equations to be used. If L < 3 Of] the“. the equations will have to be slightly modified and the corresponding equations are 0' “(3:2 —L —l , for 0.SS—z-Sl.5,andifz+0'fl >L-'—fl [4’5] ab 8 of 2 92 ' 1 “(0:2 2 _ 6} +—5_ a 8 O' 0' 3 6 ’ b H {Ir—21’ ) [4-6] 0' for 0.5s-Z—s15 andif “of, 3L...__1?'_. or 2 a(z)___3 L-Z 0f] a 8 0' 0' 3 ” 1" 3(z-—”-(L—z)) 0” [4-7] , for 1.53-5—sL——— o 2 The adsorbate-adsorbent interactions in slits are given by the partially-integrated 10-4 Lennard—Jones potential [Lee, 1987] (Equation 3-10), where the interactions are summed over the two walls. Cylindrical Pores Any molecule in a cylindrical pore is subject to a force from all sides of this cylin- der (Figure 4-2). The configuration integral in pores is given by )=}-°}i”(‘,‘2—r‘— +22) drd0dz [43] where r and z are the radial and axial distances. Using the SLD approach, the region in- side the pore is divided in three parts: 1) at the pore surface; 2) near the pore surface; and 3) far from the pore surface, and the details of these calculations are given in appendix 1. 93 Axis of symmetry &\ R i l‘ p=pa l.’ I . U‘ ‘1 ‘in”. .‘0 "V? ' “1": y." .t" ‘ 1" his,» h .t =5". 3.. ~ ! v ‘- "J. K’u-uz, ,‘ ,. .-a~ . .l 't ’>_ J' " at; atavv ~e.u tr \ 1 )’ -: al'f ‘15 ’ .2. —- A v 1 '03 :31 r ‘5: ” " 7“ "ii“? 5.9 4 : at Figure 4-2: Cylindrical Pores 94 The resulting solution leaves us with an integral to be solved for a(rl):ab, and it is the longest part in the calculation of the amount adsorbed. The results can be summarized as follows: — 62 _Zz a. 2 2+ 2 q 3 r” 14 cos'l ———fl dz- I 1t— r, Z 3,, a(fi)=3°r ° 4617 2’: "1’2z(4rI +2) 0.. 7‘ 1 , ... . 1 n ’ —— d0dz+— L 41:]0 4132c0520+z2 246},r [4-9] G orr-R—-—fl- f l 2 j” cos-'(v)dz-—j°’ j°°“( 1 d0dz 3 ° 49} (w 222+2) “(0:303 ab 1t —-l-r 1; - 0’5 -tan"(ofl./w) (19+ 1; b 4 0 4w3 2w2(w2+o}) 2W3 120} [4-10] for R— ”Sr,< —-——-fi— 363 3O “(4): 1’ “3 —1 —d0, for osn R—ji- in all cases and Equation 4-11 is not required. In cylindrical pores the adsorbate-adsorbent interactions are given by the potential of Tjatopolous et al. [1988] _ 2 ‘I’(x,Rfi)—n0;fl EM 1 [4-12] where x is the distance of closest approach between the fluid molecule and the pore sur- face, Rf; = R-t-o,/2 is the pore radius, F [ a, B; Y: 8] is the hypergeometric series with pa- rameters 0t, [3, y, and n is the number density of carbon atoms. Algorithm for solving the density profile The algorithm for solving the density profile and calculation of surface excess are given in the previous chapters. In slits and pores the only difference is the calculation of 96 a/ab, which is given by Equations 4-2 - 4-7 for slits and 4-8 - 4-11 for pores. The amount adsorbed in slits is calculated as L—ofl/Z Amt =j p(z)(S. A. )dz [4-13] non/2 and in pores as Amt = LR-” 27:7Lp(r) dr = IOR-ak r S”: p(r) dr [4-141 R _ S 2 where S.A. is the experimental surface area (for e.g. BPL activated carbon has a surface area of 988 m2/g) and L is the length of the cylinder. It was assumed that the experimental surface area was 21tL(R-o,/2). MODEL DEVELOPMENT - BINARY MIXTURES In binary mixtures, the van der Waals mixing rules are used to determine the prop- erties of the mixture [Prausnitz et al., 1986] M i=1 j=l a=izyiyjaij [4-15] b = 2m 14-161 whe COP. \ Wher 97 a,j =aj, =,/a,,afl(1—k,,) {4.17} where y,- is the mole fraction of component i, and k,-,- is the binary interaction parameter, a constant introduced to obtain better agreement in mixture equation-of-state calculations. Equation 4-1 may be written for the fugacity of each component as 22 y .14.,- ,- 3, .- lnty—p]=-;-m-m. .. 7 Where A, B and z are as discussed in chapter 3. in =1 [4-19] i=1 Algorithm to solve density profile For a binary mixture, the above equations result in solving two simultaneous equations (fugacity of each component) for both composition and density with position, subject to the constraint that the sum of the mole fractions of the two components is 1. A technique developed by Asselineau et al. [1979] for calculating VLE data has been adapted for this purpose. In this method, Equations 4-17 - 4-19 are written such that the right hand side equals zero. These equations are called the objective equations, G], G; and 6;. GI =l'9'—P—1 [4-20] fl 98 P (32 =flL—1 [421] f2 G3 = l-yl "Y2 [4’22] Partial derivatives with respect to the independent variables (pressure and local composi- tion) are then written in matrix form, which are then inverted to recalculate the objective functions such that G = \/G,2 + G: + G; is minimized. Further details of this method may be obtained from Asselineau et al. [1979], and the resulting computer program is at- tached in the appendix 2. RESULTS The diameters of the fluid, solid and solute are the Lennard-Jones parameters as tabulated in Reid et al. [1987], and are given in chapter 3. The of, is calculated as the arithmetic mean of the fluid and solid, while the fluid-solid potential parameter (cf/k) is adjusted to match the magnitude of adsorption. van Megen and Snook have done Monte-Carlo simulations of Lennard- Jones molecules (that simulate ethylene) on slit-like pores [1985] (Figure 4-3). Their simulation results cannot be quantitatively compared to either experimental or SLD model results be- cause their simulation of bulk ethylene does not represent the experimental bulk properties 0f ethylene; e. g. the simulation critical temperature (Tum) of ethylene is 226 K, as op- posed to Tc = 282.4 K seen experimentally. Figure 4-4 shows the corresponding model 99 0 5 -£ is 3 ,1-0 P Figure 4-3: Grand Canonical Ensemble Monte Carlo Simulation of Ethylene in a Slit (TIT 5°35) -h=eo, - - - 11:10.0, ----h=5.0, - ----- 11:3.5 nNEEOE ..€ mwuwvaXmH 100 30 ---h=3.5 , — - ------ h=5 .1 25 "' ----h=10 .1 I I. @20 ." .. I s ,- O . E ,/ $15 " / I (D ‘,' '” . .r. .......................... C3 ----------- .1 g 'O l".’ I.” 10 "' '.,o".’./ ”’_~7.:él".. ___________________________ 5 '25.." l 0 02 DA 00 03 PIP“at Figure 4-4: Ethylene on Carbon Slits at Tl'l'c of 085 101 predictions for adsorption of ethylene on a slit. Their calculations were done at T/T* of 0.95 (where T* = 202 K), yielding a ratio of TIT“...m of 0.85. We used the same ratio of WTC (resulting temperature was 239 K) for our calculations. To convert the excess in 1.tmol/m2 to the units used by van Megen and Snook, divide by 9.323. ‘b’ is the separation between the carbon atoms in terms of ethylene molecular diameters. Qualitatively, the SLD model shows all the trends seen in the Monte-Carlo simulations. At low slit widths the pore fills quickly, resulting in Type I isotherms. There is also a slight decrease in the excess as the pressure increases, as reported by van Megen and Snook. As the slit width is increased Type IV isotherms are seen. Figure 4-5 shows the adsorption of ethylene in pores of different diameters at low pressures. These results are compared to the data of Reich et. al. [1982], who have stud- ied the adsorption of ethylene on BPL activated carbon with a surface area of 988 m2/g. The value of the ef/k that was used in Figures 3-1 and 44 has been retained in this case. The model predictions with a pore diameter of 2.45 nm match the experimental data rather well. However, activated carbon has a pore size distribution with pores ranging from mi- cro to macropore range, with an average pore size of 3.2 nm, with largest portion of pores in the 1.5 - 2 nm region. As seen in Figures 4-6 and 4-7, at temperatures of 260.2 and 301.4 K, the model predictions at 2.45 nm are not as accurate. A factor in this tempera- ture dependency may be due to the fact that in order to truly represent the experimental data, we need to sum the amount adsorbed over this pore size distribution. At small pore widths type I isotherms are seen while type IV isotherms are seen at large pore widths. The amount adsorbed for a pore width of 4.56 nm is actually smaller than the amount ad- sorbed in a 3.29 nm pore for pressures up to 6 bar at 212. 7 K and for the entire pressure Amount (mmol/g) 102 20 - - - - Pore Dia = 4.56 nm 18 __ ----PoreDia=3.29nm _._____ Pore Dia = 2.45 nm I: " - - - Pore Dia = 1.60 nm __ ........ } .-'- ------- ‘ 16 ‘- ------ Slit Width = 1.61 nm ,-’ . :1 Experimental Data l.’ .t’ / 14 -. // , 0" I 12 "" I" l I ./ /' 0’ I . 10 "' " 1 . 1' ................. ......... S T u D 0 1 2 3 4 5 6 7 Pressure (bar) Figure 4-5: Ethylene on BPL Carbon at 212.7 K Amount (m mollg) 16 103 14- 12- O 1 r ------ Pore Dia = 4.58 nm - - - Pore Dia = 3.29 nm Pore Dia = 2.45 nm ---- Pore Dia: 1.60 nm :1 Experimental Data Figure 4-6: Ethylene on BPL Carbon at 260.2 K Pressure (bar) I 8 12 16 Amount (mmol/g) 104 - — - Pore Dia =1.60 nm 9 -.. -— - - Pore Dia = 4.56 nm -— - — Pore Dia = 3.29 nm Pore Dia = 2.45 nm :1 Experimental Data / 0 4 8 12 16 Pressure (bar) Figure 4-7: Ethylene on BPL Carbon at 301 .4 K 105 range at 260.2 K. Model predictions of adsorption of methane at high pressure were also conducted. The results obtained were qualitatitively similar to the data of Barton et. al [1984] (Figure 4-8). In Figures 4-9 and 4-10, the model predictions of adsorption of a binary mixture of ethylene and methane are shown. In this case the diameters of both ethylene and methane are assumed to be 0.4 nm to simplify calculations near the surface of the pore. The fluid- wall parameter for ethylene is retained as 140 K, while for methane it is taken as 107 K, which was calculated by fitting it to the pure gas adsorption data. A binary ethylene- methane interaction parameter of 0.022 was used [Sandler, 1989]. The composition of ethylene in the adsorbed phase is plotted as a function of pressure at 212.7 K with a bulk ethylene composition of 74%. The model predictions for the selectivity shows that the composition in the adsorbed phase is inversely proportional to the pressure, as seen by the experimental results of Reich et al. [1982]. Note that the model assumes a slit-like geome- try to model the system. In Figure 4-5, we showed that a 1.6 nm slit shows trends similar to that of a 2.45 nm pore for ethylene adsorbing on activated carbon at 212.7 K. We used the slit-like geometry to represent this binary system to show that the SLD model is capa- ble of showing the right trends with respect to composition and amount adsorbed in binary mixtures, and as seen by the figure, the SLD model shows great promise for mixture modeling. Figure 4-10 shows the model prediction for the amount adsorbed as a function of pressure. As seen by the figure the SLD model qualitatively predicts the amount ad- sorbed, and shows all the right trends seen in the figure. 1 Figure 4-11 shows the composition of ethylene in an ethylene-methane mixture at 260.2 K mixture. At this temperature, too, the SLD model predicts the selectivity of eth- 0.3 106 0.25 -- .0 M 1 Amount (9 CH4lg carbon) 0 E L. 01 0-05 ' ---,--oPore Radius 1.11 nm - — - Pore Radius 2.05 nm Average 0 . 1 1 : i o 40 80 120 160 Pressure (bar) Figure 4—8: Methane on Activated Carbon at 243.15 K Composition of Ethylene 107 0.985 1:1 :1 Experimental Data 0.98 -- —Model Prediction 3,’ a 0.975 "" .C 0. 1:1 '0 3 0.97 -- a L O 3 0935 < 0 5 .E 0.96 -- 0.955 -- 0.95 1 1 1 1 1 1 0 l 2 3 4 5 6 Pressure (bar) Figure 4-9: Adsorption of Ethylene - Methane mixture on BPL carbon at 212.7 K and Initial Bulk Ethylene Concentration of 0.74 108 Amount (mmollg) \1 (II 7 __ O 6.5 ~- 0 6 " Eth — ylene - Amt 5.5 ~~ 0 Data - Ethylene Amt. ’5 1 1 ' fi' 1 1 1 2 3 4 5 6 Pressure (bar) Figure 4-10: Adsorption of Ethylene - Methane mixture on BPL carbon at 212.7 K and Initial Bulk Ethylene Concentration of 0.74 Composition of Ethylene in the Adsorbed Phase 109 0.8 — Model Prediction 0.78 ~- :1 Experimental Data 0.76 -- 0.74 1 I r 0.72 1 0.66 - I 0.34 , 1 1 0 5 1 0 1 5 Pressure (bar) ' Figure 4-1 1: Adsorption of Ethylene - Methane mixture on BPL Carbon at 260.2 K,’and Initial Bulk Ethylene Concentration of 0.235 |||||| Ill-ll ......... 1 11 H 8 \f 11 D1 3 110 ylene very well. Figure 4-12 shows the adsorption of two members of a homologous se- ries propylene and ethylene adsorbing on activated carbon. This activated carbon is called Nuxit-AL, which is made by a Hungarian manufacturer and its characteristics could not be obtained readily, it was approximated to have the same physical properties as BPL carbon. Since ethylene and propylene are similar molecules, the binary interaction parameter was assumed to be zero. The experiments conducted by Szepesy and Illes [1963] were done by maintaining the pressure constant and varying the concentration of the bulk propylene. Here we wanted to see the model predictions over a wide range of concentrations. As seen in Figure 4-12, the model can predict the concentration in the adsorbed phase to a very high degree of accuracy. Once again, modeling the carbon as a pore, and incorporat- ing the pore-size distribution data into the model will enhance the model predictions. DISCUSSION As discussed in the two previous chapters, the limitations of the SLD model can be attributed to: 1) the local density approximation; 2) 1ack of structure in the model; 3) choice of the fluid equation-of-state to describe the fluid properties. In the previous chapter, we showed how modifying the fluid equation to incorporate the Peng-Robinson equation dramatically improved the model predictions for adsorption in flat surfaces. Here, we modify the SLD approach to predict adsorption in slits and pores. Model pre- dictions are compared to both molecular simulations and experimental data, and as seen by the results the SLD approach can quantitatively predict adsorption in activated carbon by approximating the BPL activated carbon to be a 2.45 nm pore at 212.7 K. However, as 111 0.9 :1 0.8 -- n .E ‘a’ 2 0.7 ~- 3 3 . 2 2 D. n, 0.6 ~- h 0 8 C .D g 3 0.5 -- 3 11 'E < 3 0.4 -_ g :1 U 0 3 __ —Model Prediction ' :1 Experimental Data 0.2 1 1 1 1 0 0.1 0.2 0.3 0.4 Concentration of Propylene in Bulk Phase Figure 4-12 : Adsorption of Propylene- Ethylene mixture on BPL Carbon at 293.15 K and 1.013 bar SU CV HO Car bui C165 din are 112 the temperature is increased the model predictions get worse. It is hoped that by suitably incorporating the pore-size distribution in the model i.e. by calculating the total amount adsorbed as the sum of the amount adsorbed over different pore sizes, the model predic- tions will be quantitative. Another important feature is the fact that the fluid-solid interac- tion parameter for pores was the same as that of flat walls. This suggests that the fluid- solid interaction parameter, while not a theoretically determined parameter, can be found by fitting the data at one temperature and for one surface. Once such a table can be writ- ten for a number of fluids, this model can be used as a predictive model over different surfaces. It also strengthens the universality of the SLD approach. Another limitation of the SLD model is the lack of structure in the model. How- ever, at pore sizes smaller than 3 molecular diameters, Figure 4-13 shows that the ratio a(r):a,, is symmetric about the center of the pore but the maximum is off center. As a re- sult, the calculated maximum density would be where a(r)/ab is maximum. The trends can be attributed to the fact that at pores smaller than 3 6f, there is a packing effect of the molecules such that if the particle is at the center of the pore then in the horizontal plane no other particles can be present, while if the particle is off-center, then another particle can exist in the same horizontal plane. This suggests that there is some inherent structure built in this model due to the sizes of the molecules which may lead to exclusion of parti- cles. This ‘structure’ is not seen in slit-like pores. Slits are infinitely long in two- dimensions, while pores are infinitely long only in one-dimension. Hence, more molecules are in the vicinity of a given molecule in a slit, and the packing effect is less important. 113 0.2 d- 0.15 ~- {1 \ E m 0.1 —~ 0.05 -- — Pore Dia = 1.395 nm -0.5 0 0.5 RIO” Figure 4-13: a(r)/a, for a Pore that can hold 2.5 ethylene molecules in the horizontal plane 114 The final limitation of the SLD model may be attributed to the local density ap- proximation. Figures 4-14 and 4-15 show the adsorption of ethylene on graphon and on activated carbon at 212.7 K, by treating ethylene as a homogeneous fluid (a(z)=ab.), simi- lar to the Barrer and Robins van der Waals model. By comparing these figures with Fig- ures 3-1 and 4-5, it is obvious that the volume occupied by the adsorbent must be ex- cluded. The local density approximation makes the SLD model tractable for routine calcu- lations. It can also solve the entire density profile over wide pressure ranges within a few seconds, and can give quantitative predictions for the surface excess and amount ad- sorbed. Hence, the local density approximation, while a limitation, significantly improves the model predictions over homogenous models. 1 There is also an incorrect limit in a/ab in the model for slits and pores as a slit/pore width of one molecular diameter is approached. To discuss this inconsistency let us dis- cuss slits. In such cases, for slits, the SLD model predicts that _"..=.3_[_P__1]=o atEIL=1 [4—23] This requires that as the wall separation of the slit approaches 6”, aka, approaches zero. The physical inference of this result is that either the force of attraction between the parti- cles is zero or that particles are not present. However, particles can be present in a single Plane when the slit is one molecular diameter in width and they certainly attract each. other. This suggests that this SLD methodology will not work in its present form for slit SiZes approaching 0'5. The problem arises due to the application of mean field theory that 115 45 263.15K —— 273.15K — -—293.15K - - - - 313.15K A "E = 3 :3. V m rn 8 gag fi,‘ 4 \ ° ,-—o\---~- -" 00°” \° ‘~‘ 0 ow \\ \\‘~ \\ O 60 80 100 120 Pressure (bar) Figure 4-14: Adsorption of Homogeneous Ethylene on Graphon 116 b 0'! 0 Amount (mmollg) 00 Du — Model Prediction :1 Experimental Data Pressure (bar) Figure 4-15: Adsorption of Homogeneous Ethylene on BPL Carbon at 212.7 K 117 neglects particle size, at an interface where particle size is most important. Mean field theory usually assumes that fluid particles are present in all space between the walls of a slit. At an interface this is incorrect; the reason can be visualized by considering a flat wall. In flat walls, we begin integrating from z=0.5 of, because half the volume of this particle will not be in contact with any fluid particles. This means that no fluid centers can exist between 0 and 0.5 0,7, which is consistent with particles having finite size. As we move the fixed particle away from the wall more and more of the particle surface will be in contact with other particles, and it can be surrounded by other particles at z 2 20' f. In a confined slit the SLD approach is applied for each wall, excluding particles inside 0.5 03 from each wall. When the width is just one molecular diameter, the limits overlap, thus the entire volume is excluded. One way to correct this inconsistency is to refitg/ab such that the integration be- gins from 2:0 6])“, as opposed to 2:05 of}. This would neglect particle sizes for all but the fixed particle. The problem with this solution is that ‘I’,, includes particle size, and at Z<0.5 6,; ‘19, becomes large and positive, and hence excludes particles, thus introducing a mathematical inconsistency. Another approach would be to calculate the configurational integral for a slit of one molecular width as a base reference, and add to this integral as slit widths are in- creased. Presently, however, no clear and simple solution to this problem exists, because the problem of finite particle size is simply transferred from the first layer to each subse- quent layer. 118 In order to investigate the behavior of a/ab in confined spaces approaching off, we can easily consider packing of disks in a two-dimensional channel. Solving for the con- figurational energy of closest packing in a two-dimensional channel for disks of a finite size by summing a l/r6 potential, the configurational energy for a channel of one molecular diameter width is approximately 2% lower than the configurational energy of a channel 1.5 molecular diameters wide. Based on this two-dimensional result, it seems reasonable to assume that a/ab is constant for slit widths below 1.5 off. We chose 1.5 molecular diame- ters because 1 molecular diameter represents total constraint in packing, while 2 molecular diameters represents total freedom in packing. Extension of the two-dimensional summa- tion technique to three-dimensions is most easily and realistically achieved by conducting molecular simulations in small slits, and determining dependence of configurational energy on slit size to further refine a/ab. Molecular simulations can be conducted also for pores to see if the approximation a/ab = constant is followed for pore widths < 1.5 ofl. CHAPTER 5. CONCLUSIONS There are many different models that describe adsorption ranging from the simple empirical fits to the theoretically sound molecular simulations. The simple are widely used because of their simplicity. Models based on statistical mechanics and molecular simulations are theoretically sound but significant calculation time. A model that can retain the simplicity associated with the empirical models and contains the basic physics of the adsorption problem can be very useful. This is the basis for the development of the SLD model. The SLD model is a simple model based on spatial invariance of the chemical potential, along with a cubic equation of state to describe the fluid properties. The SLD approach can model a variety of adsorption isotherms - from the low pressure region to the supercritical region. The model provides information about the density profile of the adsorbed fluid. In the subcritical region the SLD model predicts Type II and Type III isotherms on flat walls. It also predicts phase transitions at the surface (under certain conditions), as well as an abrupt transition between an adsorbed gas layer and fluid layer (again under certain conditions). Negative adsorption is sometimes predicted, similar to the predictions of Sullivan [1979]. In the supercritical region the model shows the experimentally observed cusp-like behavior near the critical point as well as the crosSovers seen at higher pressures. 119 120 Quantitative modeling, extension to Types 1, IV and V isotherms, prediction of clustering in supercritical fluids, and adsorption of multicomponent mixtures are discussed in chapters 3 and 4. By using the Peng-Robinson equation, the SLD model can quantitatively predict the adsorption of pure gases on flat walls. By modifying the solid- fluid potential to be two body fluid-fluid potential, the SLD model can predict clustering in supercritical fluids. The local density approximation calculations were modified to account for this change in geometry. Comparisons to experimental fluorescence spectra showed that the SLD model can predict the phenomenon of clustering. In slits and pores, the SLD model predicts types I, IV and V isotherms, and can also quantitatively model adsorption. Using the pore-size distributions will improve the model predictions. By approximating activated carbon to be a slit, adsorption of binary mixtures were carried out. Calculations were done for two mixtures - ethylene & methane and propylene & ethylene. The ethylene-methane mixture calculations were done at two different temperatures - 212.7 and 260.2 K. These calculations show that the SLD model can qualitiatively predict the selectivity of a binary mixture. The SLD model qualitatively predicted the amount adsorbed. By modeling the activated carbon as a pore and suitably incorporating the pore-size distributions, the model predictions can be enhanced. The SLD model serves as a good approximation and can provide rapid surveys of adsorption behavior to guide more detailed and time-consuming simulations. This. model is not intended to replace any of the more complete theories such as Monte-Carlo simulations, molecular dynamics, or density functional theory, but instead provides a simple and approximate solution to the adsorption problem, The model may be viewed 121 as a compromise or a bridge between the two-dimensional equations-of-state models, Frenkel-Halsey-Hill theory, and the more rigorous density functional or integral equation approaches. APPENDICES The van der Waals equation and configurational energy calculations - Flat Walls If we consider a canonical ensemble, the partition function Q is given by [Vera and Prausnitz, 1972] Q._1_V~(m,-_2_)”fl" N! ’ 21:1' A3 where A = h/(21r ka)”. The Helmholtz energy A and the pressure are related to the partition function: F . - 91 . u .332 3" r.» 3" rN 31an N arm 61’ 61’ where the molecular rotational, vibrational, electronic partition function, qme, is assumed to have no volume dependence. The van der Waals free volume V}, for a system with N molecules, is given by V, . V- .11. 1. NA The repulsive part of the pressure is P"? - km 1 RT V - Lb v b NA 123 If <1) is the sum of two-body interactions between an arbitrarily selected central molecule and all other molecules around it, and (b( r) is the two-body interaction potential, and g( r) the radial distribution function, then, d>(r) g(r) 4urzb VIZ ° ' 1 For a van der Waals fluid, following McQuarrie [1976], g( r) = O, for r s o, and g( r) = constant, for r > o, where o is the hard-core radius of the fluid molecule. To simplify notation, the subscript fl is omitted from 0 throughout the appendix. If we assume that the molecules interact with an infinite hard-core repulsive potential and an inverse-sixth attractive potential, then for r > o, 6 Mr) - -e,°—6 f where 6,, is the maximum energy of attraction between a pair of fluid molecules. In the bulk fluid, this leads to a” - 21: 61.03 Ali/3 124 Nonhomogeneous Systems. Consider a fluid molecule of diameter 0 at location 2 in the vicinity of a wall, where 0.50 s z s 1.50. For convenience of integration we define a cylindrical coordinate system with the origin at an arbitrary molecular center at z, and let y be a dummy variable to denote axial distance and let r denote radial distance. The configurational energy can be calculated from the two integrals: With the local density approximation: p02) z p(z) 1b - -1/2e 5 NA _.— ,o u p(z)“ 6) Since a is proportional to , for 0.50 s z s 1.50, we find 3_(Q_ em _(5 6 2) am «puma» 16 ' TE : For 1.50 s z < co, the integration is performed in three parts: 11111 the 125 With the local density approximation p(y) = p(z) these can be integrated to give: 126 and 27 RITE T a M" 64 PC TM NI.— Note that atz= 0/2,a=abu,k/2, and at z - a, a - am Note that a and its derivative with respect to z are continuous functions at z = 1.50. 127 Calculation of Configurational Integral for Clustering This calculation is similar to the configurational integral calculation for adsorption on a flat wall and further details are given in the appendix of a previous publication [Rangarajan et. al., 1995]. If O is the sum of two-body interactions between an arbitrarily selected central molecule and all other molecules around it and (p(r) is the two-body interaction potential, and g( r) is the radial distribution function, then O = I: %¢(r)g(r)4m'2dr For a van der Waals type fluid, following McQuarrie [1976], g( r) = O for r S 0, and g( r) = constant, for r > 0' , where 0 is the hard-core radius of the fluid molecule. Note that the subscript f has been dropped from 0 for the appendix to simplify the equations. If we assume that the molecules interact with an infinite hard-core repulsive potential and an inverse sixth attractive potential, then for r > 0' , O.6 ¢(r)=_efl-r_6- where e 1,. is the maximum energy of attraction between a pair of fluid molecules. In the bulk fluid, this leads to _ ~47refl. 0'3N,l ”’ 3 p 128 27: 61,. O'BNj ab = 3 The relation between the partition function, the attractive and repulsive part of the pressure, a and O are given in previously [Rangarajan et. al., 1995; Vera and Prausnitz, 1972]]. Nonhomogeneous Systems. Consider a fluid (solvent)molecule of diameter 0 at location r in the vicinity of a solute molecule of diameter 0, (typically larger than 0'), where 0.50 S r — O; S 1.50_ Using cylindrical coordinates with the origin at the center of the solute molecule, we integrate across all space with r denoting radial distance from the center of the solute molecule, x be the radial distance from the center of an arbitrary molecule (parallel to the intermolecular vector) and y be a dummy variable denoting axial distance (perpendicular to the intermolecular vector). The configurational integral can be calculated from five integrals: _ 6 °° °° 270"!de (D1 - ‘ Ear 0' NALL P0006337); ~ ~ 2nydydr __ 6 (DZ _ 617 0' NA110-10 p(r,y) (r2 +y2)3 _ 6 r—'" °'° ZWdydx ‘1’: “"517 0 ”Ala 1WP(xJ)——(x2 +y2)3 r+r' 0° ZWdydx (D4 = - EH 06NA £_r.jf——r.2_(,_x)z p(x,y) (x2 + yz )3 Znydydx O=—e 0'6N .. .. x, ———- 5 fl AJIr+r'J0 p( y) (x2 + y2 )3 where _ (0+03) I 129 r’2 + r2 —(1'2 2r’ II O=O, +O2+O3+O4+O5 With the local density approximation: p(x, y) = p( r) and since a is proportional to O, we find that a(r)_¢(r) ab (Db andleadsto 3 8 2 , a3 1 1 2 3 1 ar =—a —+— r-r +—- — +—O' () 16 b(3 O'( ) r(r’2+r2—2rr” r’2+r2+2rr’) 3 ((r+r’)3)) Similarly, for 1.50’ S r — 2.24 < oo , the integration is performed as before with an additional integral. As the fluid molecule under consideration is moved to a distance greater than 1.50 the additional integral accounts for the cohesive energy in the region between 0' I, S x S r —0. This leads to a six-part integral resulting in 3 16 203 l 03 1 1 203 1 a(r): ab( - I3+ ( I2 2 I ’2 2 I)+ I3) 16 3 3 (r-r) r r +r -2rr r +r +2rr 3 (r+r) These modified a( r) are then used in the calculation of the fluid fugacity. 130 Calculation of Configurational Integral for Slits Definitions: 2 is the distance from the surface of one of the walls. L is the distance between the two wall surfaces. r is the radial distance. x is a dummy variable that denotes axial diatance. Case 1. Particle is near the wall: 92-5 2 s 3% The integral is carried out by splitting the regions as follows: Region 1. x ranges from 0 to z-0/2 2-2 21v (D1: 4€JO6NAJO 2 Efip(x)mdrdx 1 ch =—2 3an [£-—] 1 8J0 A (Z) O 2 Note that the steps needed to carry out the integration are given in detail in the integral for flat walls. Region 2. x ranges from O to 0' <1), = -4eflo6N,j:j;;27p(x) 131 ( 22.:tr2)3 rdx x r (I)2 = 46170 3N,lrtp(z) Region 3. x ranges from 0 to L-0/2-z (D3 = ‘48306NA LL-olz-zjo p(x) O (I)3 =-—481,0‘5 2 (D3 = -38fi6 3NA1tp(Z) —28fl63NA1tp(z) C: ~ 22nr23lrdx (x +r) L-0I2—zp(x) NARI —de 6 x r - 1 1- (5-1-5f _ 0 2 0 - +2+3 5 1 —+3- L 1 3 aim-£1 _ 0 2 0 _ 132 Note that for particle near the other wall, the result is a mirror image of the above result i.e. replace 2 with L - 2. Case 11. For Large Slits, there is also a region 1.5 S S — 1.5 Cl|1~ i 0 In this case, the axial distance is split into 4 parts Region 1.x going from 0 to z-0/2 6 Z-O/Z 0- <1>1 = -4e is N, j [0 p(x)———3drdx 2m (x2 + r2) 2 (bl = _EEJO3NA1tp(z) 1" Region 2. x going from O to 0 - the result is same as case I - region 2. Region 3. x going from O to 0 - the result is same as case I - region 2. Region 4. x going from 0 to L-0/2-z - the result is same as case I - region 3. =,+2+3+4 133 (I) = -28fl.03NArrp(z) -Z—+—— Case III. In small slits i.e less than three fluid molecular diameters big, all three regions may not be present. In such cases, case I - region 2’s upper limt will become the same as case I - region 3’s upper limit i.e. region 2’s upper limit will become L—0/2-z. Region 1 will be the same as case I - region 1. . L-0/2—z . 21V O2 =—4SJG6NAJO JWP(X)WH1X . L 1 .2 z -..,.a~,.....[;-_-£] O = O, + O2 3 L (I) = -28 50’ NARP(Z)[; - I] In the limit that L = 0, the integral reduces to a two-dimensional integral. Here the ‘ solution is given by 134 «1 ,21trdr 4),, “we”, p T where p’ is given by mol/area. For slits smaller than 2 0 in width the density is a constant. To convert density in moles/volume to moles/area multiply by the slit width L. i.e. However, noting that in a slit there are two walls, i.e. the surface area is twice that of a flat Wall. the equation has to be modified to become USing this equation we get Note that at the limit L: 2 0 a; ab 16 However, the derivative is not constant at L: 2 0. Q|r~ ll colt.» II A 136 Calculation of Configurational Integral for Cylindrical Pores Definitions: The distance between the fixed particle and any particle is given by a. R is the distance from the center to the pore surface. r is the radial distance, and r, is the distance from the center of the pore and the fixed particle. 6 is the angle formed between the center of the pore, the fixed particle and the projection of any particle on the horizontal plane. 2 is the distance between any particle and the horizontal plane in which the fixed particle is located. a=\/r2 +22 The cylinder makes a circle in the horizontal plane. The locus of the circle is given by the equation where x and y are the x and y coordinates from the center of the circle (r is the radius of this circle). x = r1 — rcos(9), y = rsin(0) Solving the above two equations we get 137 r = rl cos(9)+\/(R—(—;-)2 -r,2 sin2(9) where 0' is the fluid diameter. Note that the subscript ff has been dropped from 0 to simplify notation. Now, the configurational energy is given as _ 6 r _ r <1>_—4e,,N,o Jilfidrdedz — [SH-[firdedz where B = -4eflNA0°. I. Particle away from the wall Region 1. z varies from 0 to °< 2 Fr J4. cosl)+"(R—%)2-r.2 sinze rdrdedz 0 0 B: o (r2+zz)3 138 when ~ 1 1 =J;J:< 4Z4 - 2 _2 F9 dz 4l:’i cosO+\/[R-%) ""12 sinZBJ +22 . 1 2 Let b = r1 c059 +\/(R—%) --r,2 sin29 Solving this equations yields O 1: 1.11 11 0 _tan"(0/b) 9 F 1203 '71" o 4123 -2b2(b2+02) 2b3 Region 2. z = 0 to 0 _ = cg]: J: qucose+,’(R-—)’-rfisin’0 _r___drd0dz (r +2“) Just as before solving the above equation yields 2: 1: _1r 0 +tan"(0/b) B 403 4o 2b2(b2+02) 2b3 139 Hence, for particle away from the wall i.e. O S rl S R — 30' / 2, O is the sum of the two. II. Particle nearthe wall i.e. R-30 /2Srl SR—o /2 Region 1. For 2 = 0 to °<, the integral remains the same as in 1. Region 2. For z = o to 0 In this case, 6 cannot go from O to 1t, but can only go from O to O’ r12 +02 -22 _(R__)2 cos"I 2 , because of exclusion due to the presence of the fixed ZnJoz—z particle. Note that at z=‘/02 —(R-%-rl)2, c050 =—1, 6 =1: The integral in the r direction is the same as before. This yields d) _ l w -l l cos"w 1 F— 4641.0 COS (WMZ-ZKJO (b2 +Zz)2 dedz 140 r,2 +02 -z2 -(R-‘—;-)2 where W: 2rl ’62 -22 Hence for particles near the wall i.e. R- 30 /2 S rI S R—0 l2 2=L'l£[ 1t 0' tan"(0/b) 0+—1—-J‘:cos"(w)dz-%I:J’:os- 13 121:3 4 m—2b2(b2+02)- 2b3 46‘ III. Particle at the wall In this case 6 varies only from O to 1t. Region 1. z varies from 0 to °< Here, rl c089 +\/(R-%)2 - r,2 sin2 9 = 2rl cosO Hence, the integral becomes 1 (b2 +22)2 141 ’2 Moose _r__drd9dz B :41]: 1" (r +z 23) O ~ 12 l 12 1 -- = —-d9dz - l dad [3 L1: 424 0": 4[4r,2 c0529 +22]2 2 2_ 1r - 1t(2r,2+zz) .10 223(4’12+ZZ)3/2 Z [V2403- Region 2. For 2 = O to 0 Once again 6 cannot go from O to 1t, but, can only go from O to 0' 2 2 2 2 cos"1 2 J 2 2 2i , because of exclusion due to the presence of the fixed r1 0 — 2 particle. 112:”1 +6 -z2—(R——) 62-12 21', JO 2 - z 2’1 Doing the calculations as before, ch 1 _ ch-z2 1 - cos-'w 1 «B— = J: COS '(T)dz -Z‘[" L d6dz 4 [4r]2 cos2 6 + 22]2 142 1r ‘1. 1r(2r,2+zz) d -l _.0;J “40 O B _ 240.3 °223(4r,2 +zz)3/2 —:Z)d —_'10 J11“ “[4rl cos 1‘9+z 1d9dz Note that in all cases 2— = _a_ = (b O, a, —l6efl03NA1tp/3 0000000000 00000000000 A ndix2- rams FLAT WALL - Pure Gas Program to calculate surface excess or adsorption isotherms The program uses Peng-Robinson equation of state, with a modified a to account for exclusion. The strength of the adsorption potential and its dependence on position is given by Psi. The fluid-fluid potential uff = 11 total - u ads. From this potential the various fluid properties are calculated. The variables used are defined below. All units used are SI unless otherwise stated. COMMON R,TC,PC,OMEGA Define Psi as a statement function Use a 10-4 potential where PSI is the negative of the intermolecular potential. SIGFW is sigma fluid-wall, in Angstroms. EPSFW is the fluid-wall potential in K ALPHA is the ratio of the spacing of the graphite basal planes (3.35 A) to SIGFW. The number density of C atoms in graphite is 0.382 atoms/A42. The equation used is that suggested by Lee (1.5). PSI(ETA)=4.0*3.1415926*O.382*SIGFW*SIGFW*EPSFW*(-0.2 OD OOOOOOO l/ETA* *10+O.5/ETA**4+0.5/(ETA+ALPHA)**4+0.5/(ETA+2.0* *ALPHA)* *4-1-0.5/(ETA+3.0*ALPHA)**4+0.5/(ETA+4.0*ALPHA) **#4) - WRITE(*,*)'Temp Plim Epsilon Sigff Sigww Tc Pc omega' READ(*,*) T,PLIM,EPSFW,SIGFF,SIGWW,TC,PC,OMEGA WRITE(*,*) 'EPSILON‘ READ(*,*) EPSFW T=212.7 PLIM=7.E5 SIGFF=4.22 SIGWW=3.4 TC=282.4 PC=50.4e5 OMEGA=0.089 OPEN(UNIT=7,FILE='denpro.dat',STATUS='UNKNOWN') OPEN(UNIT=8,FILE='adsorp.dat’,STATUS='UNKNOWN‘) R = Gas constant J/K/mol Tc = Critical Temperature K Pc = Critical Pressure Nlm2 OMEGA = Acentric Factor SIGMA = Sigma fluid-wall in m. T = Temperature K FLIM = Bulk fugacity limit. 143 0000000 010 M0 144 EPSFW = Fluid-wall potential well depth in K SIGFW = Sigma fluid-wall in Angstroms ALPHA = Spacing of graphite planes / Sigfw SIGFF = Sigma fluid-fluid in Angstroms SlGWW = Sigma wall-wall in Angstroms AB = Bulk Pong-Robinson 'a' BB = Bulk Peng-Robinson 'b' R=8.3 14 POMEG=0.37464+1 .54226‘OMEGA-0.26992*(OMEGA "2) TR = T/TC SIGFW=(SlGWW+SlGFF)/2. SlGMA=SlGFW*IE-10 ALPHA=3.35/SIGFW ALPHA = Spacing of graphite planes / Sigfw SlGFF = Sigma fluid-fluid in Angsuoms Calculate a bulk and b AB=0.45724*(R‘TC'(1+FOMEG*(1-SQRT(TR))))“2/PC BB=0.0‘T780"R‘TC/PC Write parameters to files PCB=PC/l.OES IF(TR.GE.1) GO TO 2 CALL FSAT(AB,BB,T,FUGS,PS,VV,VL) DELP=PLIMI40. Loop for bulk fugacity PB=0.0ES 1:0 FOR EACH BULK FUGACITY VALUE PB=PB+DELP PB=PLHW B=BB A=AB First calculate bulk density (DENB) and bulk pressure (BP) CALL BVCAL(A,B,T,PB,PS,V.FB) CALL PV(A,B,T,V,P) DENB=1 .OIV BP=P 145 C WRITE(7,*) PRESSURE. P C WRITE(7,")'BULK DENSITY', DENB DELETA=0.1 ETA=.9 EXCESS=0.0 K=0 ETA=ETA+DELETA 0‘ Calculate local fugacity Use the following formula to get the local fugacity (F), for a given position ETA. F=FB*EXP(PSI(ETA)/l‘) 0000 ALNFB=ALOG(FB) ALNF=ALNFB+PSI(ETA)/l" PSIV=PSI(ETA)/l' C Calculate local 11 CALL ACALCCETA,AB,A.SlGWW,SIGFW,SIGFF) C Using local parameters calculate V and local density DENL CALL VCALC(A,B,T,ALNF,BP,V) DENL=1.0/V EXCESS=SIGMA*(DENL-DENB)*DELETA*l.E6+EXCESS C DENLG is the density in gmoles/cc DENLG=DENU1 .0136 WRITE (7,102)ETA.DENLG WRWBO*) 'A'O.O'DAO.O'O'B'D',"BO'Q'DETA K=K+l IF(K.LT.200) GOTO 6 9O CONTINUE BP=P/l .0E5 DENBG=DEN B/ 1.136 WR1TE(6,*) BP, ',', EXCESS, ',', DENBG WRITE(8,*) BP, ',', EXCESS, ',', DENBG J=J+l [F (J .LT.40) GOTO 5 146 100 CONTINUE C WRITE(*,*)'Press l to continue' C READ(*,*) INT C [F (INTEQJ) GOTO 566 CONTINUE 102 FORMAT(1X,F8.2,2X.F12.5) END C?Itfiifififit.‘fittfitltfififitfitt.¥¥¥¥fittli***.**tt*#*ti¥ll¢¥¥$*********¥ C234567 C SUBROUTINES Cttttttttttttttittttttttttttttttttt*ttttutti!tttttttttttttttttttttt SUBROUTINE VCALC(A,B,T,ALNF,BP,V) COMMON R.TC.PC.OMEGA IV1=1 1V3=1 DPDV1=1 DPDV3=1 P1=O P3=0 V=0 C 1F(BP.GT .PS) GOTO 16 CALL VICALC(A,B,T,ALNF,V1,IV1) IFGVI.BQ.1)GOTO 10 CALL DPDV(A,B,T,V1,IDPDV1) IFODPDVIJSQJ) GOTO 10 CALL PV(A.B,T,V1,PI) 10 CONTINUE CALL V3CALC(A.B,T,ALNF,V3,IV3) 1F (IV3.EQ.1) GOTO 14 147 CALL DPDV(A,B,T,V3,IDPDV3) IF(IDPDV3 .EQ.0) THEN CALL PV(A,B,T,V3 .P3) ELSE CONTINUE ENDIF l4 CONTINUE IF (P1.GT.P3) THEN =Vl ELSE =V3 ENDIF CONTINUE C WRITE(2,")P1,P3 GOTO 15 C 16 CALL V3CALC(A,B,T,ALNF,V3,IV3) C IF(IV3.EQ.O) GOTO 21 C 16 CALL VICALC(A,B,T,ALNF,V1,IV1) C V=Vl C GOTO 15 C 21 CALL PV(A,B,T,V3,P3) C V=V3 15 RETURN END J‘AAJALJ.‘.LJ-AAAA-lAAALJJAJAAAAAJJJAJJJJAJJJJAAJA¢44al_JJ_JJJJJ vsTTrTTVTVTTTTTTTTTTCTTTTTVTTVVTTTTTTTTTTTWTvvvveuvlvvvlt***** 148 SUBROUTINE DPDV(A,B,T,V,1DPDV) COMMON R Input A,B.T,V and Output = IDPDV THIS SUBROUTINE EVALUATES THE DERIVATIVE dP/dv AND RETURNS A VALUE OF 1 FOR IDPDV IF THE SLOPE IS POSITIVE. i.e. THE ROOT IS UNSTABLE. 000 0 IDPDV=0 DP=-R"T/(V-B)/(V-B)+2.0"A*(V+B)/((V*V+2*B*V-B *B)**2) IF (DP.GT.0) IDPDV=1 RETURN END C.t.‘..*.***..*UUtfittfitfittfiltlfififitlt*Ililt##3##*¥*******¥*******¥ SUBROUTINE VICALC(A,B,T,ALNF,V1,1V1) COMMON R,TC,PC,OMEGA IV1=O C This subroutine uses a successive substituion method C to get the small root Vl. C Inputs A,B,T,ALNF and Output V1, IV] C Starting guess for calculating v Vl=l.05"B C Iterate 60 times D020 I = 1,60 ARG=(V1-B)/R/T IF((V1.LE.B).OR.(ARG.LE.0.0)) THEN IV 1 =1 C WRITE(9,")'INV1CAL V1: ',V1,' 3 = ',B,'ARG = ',ARG C I,'I =',I,' ALNF,ALNF GOTO 30 ELSE DENO=ALNF+ALOG(ARG)+A*V1/R/r/(V1 '"2+2"'B *V l -B .”2) DENO=DENO+AI2.82/Rfr/B’ALOG((V1+2.414*B)/(V1-0.4 14*B)) VI=B+BIDENO ENDIF C WRI'I'E(6,"')'V1,B',V1,B C STOP 149 20 CONTINUE ALNFC=B/(vr-B)-A*v1/Rrr/(VI*vr+2*B*v1-B*B) ALNFC=ALNFC+ALOG(R'T/(Vl-B)) ALNFC=ALNFC+A/2.82/B/RII“ALOG((Vl-O.414*B)/(V1+2.414*B)) ERROR=EXP(ALNFC-ALNF)-l .0 IF (ABS(ERROR).GE.0.001) 1V1 =1 30 CONTINUE C WRITE(9,“)'VI ',ALNFCALNFN l RETURN END COIOOOOQIOOOOIOIOOOttfittttfitllfiltllfifitttttttttt***¥¢**#****#****13* C234567 00 SUBROUTINE V3CALC(A,B,T,ALNF,V3,IV3) COMMON R,TC.PC.OMEGA IV3=O STARTING GUESS FOR V3 = RT/F ALNV3=ALOG(R"I')-ALNF V3=EXP(ALNV 3) ITERATE TILL DONE OR 40 TIMES DO 201:1,40 IF (V 3.LE.B) THEN WRITE(9,")'V3 LE 8', V3,‘ B ',B IV3=I GOTO 30 ENDIF ARG=B/(V3-B)-A*V3/R/T/(V3"2+2*B‘V3-B“2) IF(ARG.GT.40.) THEN D3=1./V 3 WRITE(9.‘)'V3 = ',V3,‘ ARG LARGE = ',ARG,'D3 =',D3 lV3=1 GOTO 30 ELSE DENO=A/2.824/Rfl'/B I"ALOG((V3-0.4 14*B)/(V3+2.4 14 *B)) ALNV33=ALOG(R‘T)-ALNF+ARG+DENO V3=B+EXP(ALNV3B) ENDIF 20 CONTINUE ALNFC=B/(V3-B)vA"V3/R/r/(V3‘V3+2*B *V3-B *B) ALNFC=ALNFC+ALOG(R*T/(V3-B)) 150 ALNFC=ALNFC+A/2.824 I"ALOG((V3-0.4 l4 I"B)/(V3~l-2.4 l4*B))/B/R/I‘ C warmer-y 1N v3. 1n fc, v3 ',ALNFC,V3 ERROR=EXP(ALNFC-ALNF)-l.0 IF(ABS(ERROR).GE.0.001) 1v3=1 30 CONTINUE C WRITE(4,"')'V3',ALNFC,ALNF,V3 RETURN END Ctitfitfitfifitfitfiitttitttltlfiltitifitilttittttittttittii*1!**********¥ C234567 SUBROUTINE PV(A,B,T,V,P) COMMON R C This subroutine returns the value of P given a,b,T,v C using the van der Waals equation of state. P=R*T/(V-B)-A/(V*V+2"B*V-B‘8) RETURN END C.ititttfittllfitlttltlfifilltlfitt.ttfifiliiltttttlttilttttlttt C234567 SUBROUTINE ARANGE(R1,RZ.R3) C PROGRAM TO PUT 3 NUMBERS IN DESCENDING ORDER DO 20 J=1,3 IF(R2.GT.R1) THEN TEMP=R 1 R1 =R2 R2=TEMP ENDIF 1F(R3.GT.R2) THEN TEMERZ R2=R3 R3=TEMP ENDIF 20 CONTINUE RETURN END Ct.tit.*lttttittit...ttitfittlfiififilfitltiltttltttfittttttltit********************* C234567 SUBROUTINE ACALC(EI'A,AB,A,SIGWW,SIGFW,SIGFF) C This subroutine calculates the van der Waals a term C after taking into account the effect of exclusion. 151 AB = Value of a in the bulk ' ETA = reduced distance from the center of wall The main program sends a reduced distance ETA from the center of the wall molecule, which is the basis for the integrated 9-3 potential. However the integrations for configurational energy have been done from the edge of the wall molecule. Therefore it is necessary to translate the coordinate. BETA is the distance from the edge of the wall in reduced units 0000000000 BETA =( ETA - (0.5 *SIGWW/SIGFW))*(SIGFW/S IGFF) IF (BETA.LE.1.5) T1 IEN A=AB*(5.0+6.0*BETA)/ 16.0 ELSE A=AB"'(1 .0-1 .0/8.0/(BETA-0.5)“3) EN DIF RETURN END (:t******ttttttttttttttttttttttitittttfitittittit$ttttttttttt &7l 2 3 4 5 6 72 O.‘ttttlttIttitfitttlltfitlfiO...fit.*ttttt'*tlfififitfititlttfiifit****t*** " SUBROUTINE CUBIC " I‘********¢$tttttfillltltttltttit.##1##.#I************************* " THIS SUBROUTINE FINDS THE ROOTS OF A CUBIC EQUATION OF THE "' FORM X“3 + A2‘X“2 + AI‘X + A0 = 0 ANALYTICALLY. * *ttfififitttttttttfi*tlt*‘fittttttt*lfiitliitttt*ttlttttt******t******** "' VARIABLES SUBROUTINE CUBIC(A2,A l,A0,Rl .RZ,RB,C1,C2,C3,IFLAG) DOUBLE PRECISION CHECK,DAO,DA1.DA2,PI,P2,Q,R,SS1,552 COMPLEX ES 1 ,ES2,S l .8221 ,ZZZ3,CCHECK DAO = DBLE(AO) DAI = DBLE(AI) DA2 = DBLE(A2) Q = DA1/3.D00 - DA2*DA2/9.DOO R = (DAI *DA2-3.D00*DAO)/6.D00 - (DA2/3.D00)**3 CHECK = Q"3 + R‘R IF(CHECK.GT.0.0) THEN IFLAG = 1 P1 = R+DSQRT(CHECK) P2 = R-DSQRT(CHECK) IF(PI .LT.0.0) THEN $81 = -DEXP((DLOG(-1.D00*PI))/3.D00) ELSE $81 = DEXP((DLOG(PI))/3.DOO) ENDIF IF(PZ.LT.0.0) THEN SS2 = -DEXP((DLOG(-I.D00*P2))/3.DOO) 00000000 152 ELSE 552 = DEXP((DLOG(P2))/3.DOO) ENDIF R1 = $51 + $52 -DA2/3.D00 R2 = -(SSI+552)-DA2/3.D00 R3 = R2 C l = 0.0 C2 = (SQRT(3.))"(551-552)/2.D00 C3 = - C2 ELSE IF (CHECK.LT.0.0) THEN IFLAG = 3 RR = 1."‘R ‘V RECK =1 .*CHECI( CCHECK = CMPLXCRECK,0.0) E51 = CLOG(RR+CSQRT(CCHECK))/3. E52 = CLOG(RR-CSQRT(CCHECK))/3. 51 = CEXP(ES 1) 52 = CEXP(ESZ) 21 = (51-1-52) - A2/3 22 = ~(51+52)/2 - A2/3 +(CMPLX(0.0,3“.5))"(S1-52)/2 23 = -(51+52)/2 - A2/3 - (CMPLX(0.0,3".5))"(S1-52)/2 R1 = REAL(ZI) R2 = REAL(Z2) R3 = REAL(Z3) C1 = 0.0 C2 = C1 C3 = C l ELSE [FLAG = 2 1F(R.LT.0.0) THEN 551 = -DEXP((DLOG(- l .DOO‘R))/3.D00) ELSE IF(R.EQ.0.0) THEN 551 = 0.0 ELSE 551 = DEXP((DLOG(R))/3.D00) ENDIF 552 = 551 R1=551+SS2-DA2/3.D00 R2 = -(551+SS2)/2-DA2/3.D00 R3 = R2 C I = 0.0 C2 = C1 C3 = C2 ENDIF RETURN END C C 153 PROGRAM FOR CLUSTERING ittfififiiifittttfitfititltttttt*t.‘ttttlttttttttttttttt*ttttlfllttttttttfil C234567 0000000000 0000000000000 00000000000000 Program to calculate surface excess or adsorption isotherms The program uses Peng-Robinson equation of state, with a modified a to account for exclusion. The strength of the adsorption potential and its dependence on position is given by Psi. The fluid-fluid potential uff = u tOtal - u ads. From this potential the various fluid properties are calculated. The variables used are defined below. All units used are 81 unless otherwise stated. COMMON R.TC.PC.OMEGA Define Psi as a statement function Use a 10-4 potential where P81 is the negative of the intermolecular potential. SIGFW is sigma fluid-wall, in Angstroms. EPSFW is the fluid-wall potential in K ALPHA is the ratio of the spacing of the graphite basal planes (3.35 A) to SIGFW. The number density of C atoms in graphite is 0.382 atoms/AAZ. The equation used is that suggested by Lee (1.5). P51(ETA)=4.‘EPSGS‘(1./ETA“6ol JETA"12) R = Gas constant I/K/mol Tc = Critical Temperature K Pc = Critical Pressure N/m2 OMEGA = Acentric Factor SIGMA = Sigma fluid-wall in m. T = Temperature K FLIM = Bulk fugacity limit. EPSGS = Fluid-wall potential well depth in K SIGFW = Sigma fluid-wall in Angstroms ALPHA = Spacing of graphite planes / Sigfw SIGFF = Sigma fluid-fluid in Angstroms SIGWW = Sigma wall-wall in Angstroms AB = Bulk Peng-Robinson 'a' BB = Bulk Peng-Robinson 'b' WRITE(*,")' Plim ,T,EPSILON' READ("',") PLIM,T.EPSGS R=8.3 14 TC=282.4 PC=50.4E5 OMEGA=0.089 SIGFF=4.22 154 SIGWW=7. 14 SIGFW=(SIGWW+S IGFF)/2. SIGMA=SIGFW"1E-10 Write parameters to files OPEN(UNIT=7,FILE='denpro.dat',STATUS='UNKNOWN') OPENCUNIT=8,FILk‘adsorp.dat’,STATUS='UNKNOWN') WRITE(7,*) ' Ethylene at ',T,‘ K with an epsilon of ',EPSGS WRITE(8,"') ' Ethylene at ‘,T,‘ K with an epsilon of ',EPSGS WRITE(8,"‘) 'P (Pa) Excess (micro mol/m"2) Den (mol/cc)' WRITE(8,*) 'Clus (# molecules solvent/molecule solute)’ WRITE(8,997) WRITE(8,") WRITE(8,998) 997 FORMAT(7X,'P',9X,'EXCES5',8X,'CLUS',9X.'DEN') 998 FORMAT(7X,'0,',10X,'0,',13X,'0,',13X,'0') v.0 0H0f5 TR=T/T C FOMEG=0.37464+1 .54226*OMEGA'0.26992*(OMEGA"2) Calculate a bulk and b AB=0.45724‘(R‘TC’(I+FOMEG*(I-SQRT(TR))))"2/PC BB=0.07780’R‘TCIPC PCB=PCI 1 DE IFCI'R.GE.1) GOTO 2 CALL FSAT(AB,BB.T,FUGS.PS.W,VL) DELP=PLIMI40. Loop for bulk fugacity PB=0.0E5 I=0 INN!EAKELBULKIRKENCHWFVALUE PB=PB+DELP PB=PLHH B=BB A=AB First calculate bulk density (DENB) and bulk pressure (BP) CALL BVCAL(A,B,T,PB,PS,V,FB) CALL PV(A,B,T,V,P) DENBN=1.0/V In this method, we calc. the bulk density from the program in which eta starts at 20.9 and comes inward. Sometimes do check DENBN and DENB and make sure WRITE?!) 'BULK DENSITY = ?' READ("',"') DENB BP=P WRITE(7,"')'P = ', P,’ Pa',‘ DEN CALC = ',DENBN,’ gmol/m3' WRITE(7,')'DENB = ',DENB,’ gmol/m3',’ FB = ',FB,' Pa' 155 WRITE(7,") 'DEN (gmol/m3) PMV (cc/gmol)’ WRITE(7,“) 'CLUS (it excess solvent moles/moles. solute)‘ WRITE(7,“) 'AMOUNT (# total solvent molec/molec. solute)‘ WRITE(7,"') WRITE(7,999) 999 FORMAT(7X,'ETA',8X.'DEN',10X,'PMV',9X,'CLUS',7X,'AMOUNT) C O 0000 C Start iterating in ETA DELETA=0.01 ETA=.99 EXCESS=0.0 CLUS=0.0 K=0 ETA=ETA+DELETA Calculate local fugacity Use the following formula to get the local fugacity (F). for a given position ETA. F=FB*EXP(PSI(ETA)/I') ALNFB=ALOG(FB) ALNF=ALNFB+PSI(ETA)/T PSIV=PS 1(ET A)/T Calculate local a CALL ACALC(ETA.AB,A,SIGWW.SIGFW,SIGFF) IF(TR.GE.1.) GO TO 20 CALL FSAT(A,B,T.FUGS,PS,VV,VL) Using local parameters calculate V and local density DENL 20 CALL VCALC(A,B,T,ALNF,BP,PS.V) 00000 C DENL=1 .O/V IF((DENUDENB.GT.0.999).AND.(DENL/DENB.LT.1.001)) THEN DENL=DENB ' ENDIF CLUS=SIGMA"3’(DENL-DENB)“4.'3.1416‘ETA‘ETA‘DELETA*6.023E23+CLUS IF(K .EQ.0) THEN CLUS=CLUS-DENB*4./3*3.I416*SIGMA"3*6.023E23 write(",") eta,clus ENDIF AMT is the excess amount (moles) of solvent moles in the system TV is the vol of the system at any given eta EV is the equivalent volume in the pure system TW is the vol of the solvent molecules in the actual system provided they are present at bulk density AMT=4."'3.1416"SIGMA"3‘ETA‘ETA*(DENL-DENB)*DELETA+AMT TW=4.I3.*3. 1416‘SIGMA"3"(ET A"3-1 .) TV=4./3."3.1416*(ETA’SIGMA)"3 EV=A MT/DENB+TW PMV=TV-EV PM V=PMV *6.023E29 AMOUNT is the total # of solvent molecules in the system 156 C TN is the total it of molecules in the pure system at any eta C EXCESS is the excess # of solvent molecules in the system AMOUNT =4.*3.14 l6*5IGMA**3‘ETA'ETA*DENL‘DELETA*6.023E23+AMOUNT TN=DENB"6.023E23‘TV EXCESS=AMOUNT-TN C DENLG is the density in gmoles/cc DENLG=DENUI .OE6 IF((K.EQ.1 14).OR.(K.EQ. 190)) THEN WRITE(".*) ET A,AMOUNT EN DIF WRITE (7,")ETA,',',DENL,',',PMV,',',CLUS,',',AMOU NT =K+1 IF(K.LT.2000) GOTO 6 90 CONTINUE BP=P/l .OES DEN BG=DENBI I .56 WRITE(8.’)BP.',',EXCESS,',',CLUS,',',DENBG WRITE(6,*) BPEXCESS,CLUS.PMV I=I+I C IF (J.LT.40) GOTO S 100 CONTINUE 102 FORMAT(1X.F8.2,2X.F 12.5.2X,FIO.5) END C***¢*****fitttttttttfiIt.*tlttifittfittlttittttfl***¥************tI“!!! C234567 C SUBROUTINES Cttaste-teammat-tantrum*tttstseasesetetsasteeeeatatttatetetsetta:tum C Use subroutine VCALC as per fiat wall program C Use subroutine DPDV as per flat wall program Use subroutine VICALC as per fiat wall program Use subroutine V3CALC as per fiat wall program Use subroutine PV as per fiat wall program Use submarine ARANGE as per fiat wall program SUBROUTINE ACALC(ETA,AB,A,SIGWW,SIGFW,SIGFF) This subroutine calculates the van der Waals a term after taking into account the effect of exclusion. AB = Value of a in the bulk C C C C C C C C C ETA = reduced distance from the center of wall 157 The main program sends a reduced distance ETA from the center of the wall molecule, which is the basis for the integrated 9-3 potential. However the integrations for configurational energy have been done from the edge of the wall molecule. Therefore it is necessary to translate the coordinate. BETA is the distance from the edge of the wall in reduced units BETA = (ET A*(SIGFF+SIGWW)-SIGWW)/(2.'SIGFF) 00000000 SIGMA1=SIGFF SIGMA2=SIGWW R=(SIGMA1+SIGMA2)/Z. Z=ETA*(SIGMAI+SIGMA2)/2. R1=(R*R-SIGMA1*SIGMA1+Z"Z)/(2"Z) IF (BETA.LE.I .5) THEN Al=8./3.+2."'(Z-RI)/SIGMAI A2=SIGMAI *rs/(zrmemz-z-zrm *2» A3=(SIGMAI"3)/(Z*(R*R+Z*Z+2.*R*Z)) A4=2.*SIGMAI"3/(3."(Z+R)"*3) A=AB"(AI+A2-A3+A4)"3116. ELSE A l=16J3.-2."SIGMAI"3/(3.*(Z-R)“3) A2=2.*SIGMA1”3/(3.*(Z+R)"3) A3=SIGMA1 I""‘3/(Z"‘(R"'R+Z"Z-2."Z"'R)) A4=SIGMA1“3/(Z*(R"'R+Z"Z+2.*Z"R)) A=AB"(AI+A2+A3-A4)"3./16. EN DIF RETURN END C Use subroutine CUBIC as per fiat wall program C C 158 SLITS - PURE GAS Ilfifitttttttttttltlil*i*****$***¥**#******************INN!!!********* C234567 nonooonoonoon 0000000000 0000000000000 Program to calculate surface excess or adsorption isotherms The program uses Peng-Robinson equation of state, with a modified a to account for exclusion. The strength of the adsorption potential and its dependence on position is given by Psi. The fluid-fluid potential uff = u total - u ads. From this potential the various fluid properties are calculated. The variables used are defined below. All units used are SI unless otherwise stated. COMMON R,TC,PC,OMEGA Define Psi as a statement function Use a 10-4 potential where P81 is the negative of the intermolecular potential. SIGFW is sigma fluid-wall, in Angstroms. EPSFW is the fluid-wall potential in K ALPHA is the ratio of the spacing of the graphite basal planes (3.35 A) to SIGFW. The number density of C atoms in graphite is 0.382 atoms/AAZ. The equation used is that suggested by Lee (1.5). P511(ETA)=4.0"'3.1415926*0.382*SIGFW‘SIGFW'EPSFW*(-O.2 I/EI‘A“ lO+0.5/ETA"4+0.5/(ETA+ALPHA)" l"4«l'0.5/(ETA‘1-2.0"I *ALPHA)"4+0.5/(ETA+3.0"ALPHA)"*4+0.5/(ETA+4.0‘ALPHA) ttt4) PSI2(XI)=4.0*3.I415926‘0.382*SIGFW*5IGFW‘EPSFW*(-O.2 l/XI"10+0.5/X1"4+0.5/(X1+ALPHA)"4+0.5/(X1+2.0"' ‘ALPHA)'*4+0.5/(XI+3.0*ALPHA)**4+0.5/(X1+4.0‘ALPHA) *tt4) PSI(ET A)=PSI 1 (ET A)+PSI2(XI) R = Gas constant J/K/mol Tc = Critical Temperature K Pc = Critical Pressure N/m2 OMEGA = Acentric Factor SIGMA = Sigma fluid-wall in m. T = Temperature K FLIM = Bulk fugacity limit. EPSFW = Fluid-wall potential well depth in K SIGFW = Sigma fluid-wall in Angstroms ALPHA = Spacing of graphite planes / Sigfw SIGFF = Sigma fluid-fluid in Angstroms SIGWW = Sigma wall-wall in Angstroms AB = Bulk Peng-Robinson 'a' 159 C BB = Bulk Pang-Robinson 'b' C ALPHA = Spacing of graphite planes / Sigfw C SIGFF = Sigma fluid-fluid in Angstroms WRITE(‘,*)' Plim ,TEPSILON, WIDTH' READ(",*) PLIM,T,EPSFW.ZET A R=8.3 l4 TC=282.4 =50.4E5 OMEGA=0.089 SIGFF=4.22 SIGWW=3.4 SIGFW=(SIGWW+S 1GFF)/2. SIGMA=SIGFW*1E-10 ALPHA=3.35/51GFW C Write parameters to files OPEN(UNIT=7,FILE='denpro.dat',STATUS='UNKNOWN') OPEN (U NIT =8,FILE='adsorp.dat'.5TATUS='UNK NOWN' ) WRITE(7,"') ' Ethylene at ',T,’ K with an epsilon of ',EPSFW WRITE(8,") ' Ethylene at ',T,’ K with an epsilon of ',EPSFW WRITE(7,*) ' with slit width of ',ZETA .' ethylene mol dia' WRITE(8,"') ' with slit width of ‘,ZETA ,' ethylene mol dia' WRITE(8,"‘) WRITE(8,*)'P bar Exc um/m2 Amt mmol/g Amtl mmol/g Den gmol/cc' Wimsf.) 0’" '9 09" '90," '9 0 TR=TIT C POMEG=Q37464+I .54226*O MEGA 41.26992‘(O MEGA l”2) C Calculate a bulk and b AB=0.45724*(R'TC*(I+FOMEG‘(I-SQRT(TR))))"2/PC BB=0.07780*R"'TC/PC PCB=PCII .OES IF(TR.GE.1) GOTO 2 CALL FSAT(AB,BB,T,FUGS.PS,VV,VL) 2 DELP=PLIMI40. C Loop for bulk fugacity PB=0.0ES 1:0 C FOR EACH BULK FUGACITY VALUE C PB=PLIM 3 PB=PB+DELP B=BB A=AB C First calculate bulk density (DENB) and bulk pressure (BP) CALL BVCAL(A,B,T,PB,PS,V,FB) 0000 C C C 160 CALL PV(A,B,T,V,P) DENB=1.0/V BP=P WRITE(7,"‘)'P = ', P,‘ bar' WRITE(7.")'DENB = ',DENB,‘ gmol/m3',‘ FB = ',FB.‘ bar' WRITE(7.*) WRITE(‘7,")' ETA',', ',' DEN gmol/m3',', ','Log(Fugacity)' BET A=0.5-(ZET A-1 .)l 100. DELET A=(ZE'TA-1.)/ 100. EXCESS=0.0 AMOUNT=0.0 AMOUNT1=0.0 L20 Start iterating in ETA BET A=BETA+DELETA ETA=(BETA*2.*SIGFF+SIGWW)/(51GFF+SIGWW) XI=(ZETA“SIGFF+SIGWW/‘2.-BETA*SIGFF)/SIGFW Calculate local fugacity Use the following formula to get the local fugacity (F), for a given position ETA. F=FB*EXP(PSI(E'TA)/I') ALNFB=ALOG(FB) ALNF=ALNFB+PSI(ETA)/I‘ PSIV=PSI(ETA)/I' Calculate local a CALL ACALC(BETA,AB,A.ZETA) IF(TR.GE.1.) GO TO 20 CALL FSAT(A,B,T.FUGS,PS,VV,VL) Using local parameters calculate V and local density DENL 20 CALL VCALC(A.B,T,ALNF,BP,PS,V) DEN I: 1 .ON EXCESS=5 IGMA‘(DENL-DENB)*DELETA" 1 .E6/2.+EXCESS Amount is den(gmol/m"3)"sigma(m)‘deleta'S.A.(m"2/g) to convert Amount to mmol/g multiply by 1000 AMOUNT=DENL‘SIGMA*DELETA*988."'I.E3+AMOUNT AMOUNT1=DENL‘S IGMA‘DELETA*988.I2."1 .E3+AMOUNT1 DENLG is the density in gmoles/cc DENLG=DENU1 .0156 WRITE (7,"')BETA,',',DENLG,','.ALNF L=L+I IF(L.LT.101) GOTO 5 BP=P/l .OES DEN BG=DENBI1 .E6 161 WRITE(8,*)BP,',',EXCESS.','.AMOUNT,',',AMOUNTI ,‘,'.DEN BG WRITE(6,") BP,EXCESS .AMOUNT,AMOUNT1 I=I+l IF(I.LT.40) GOTO 3 102 FORMAT(1X,F8.2,2X.F 12.5,2X,F10.S) END Cflt$ttititttltttltittt##3##tttttttttttttfi*******tilt************** C234567 C SUBROUTINES Ctttttttttttttttttttttttttttttttitttttttttttrtt*IrrlrtrtrrtntrhkhtrrlrrlrIt:Irwin": C Use subroutine VCALC as per flat wall program Use subroutine DPDV as per flat wall program Use subroutine VICALC as per fiat wall program Use subroutine V3CALC as per fiat wall program C C C C Use subroutine PV as per fiat wall program C Use subroutine ARANGE as per fiat wall program C234567 SUBROUTINE ACALC(BETA,AB,A,ZET A) This subroutine calculates the van der Waals a term after taking into account the effect of exclusion. AB = Value of a in the bulk ETA = reduced distance from the center of wall The main program sends a reduced distance ETA from the center of the wall molecule, which is the basis for the integrated 9-3 potential. However the integrations for configurational energy have been done from the edge of the wall molecule. Therefore it is necessary to translate the coordinate. BETA is the distance from the edge of the wall in reduced units SA=ZETA-0.5-BETA IF (BET A.LE.1.5) THEN A I=BETA+0.5 A2: I .- l .ISA"3 A3=1./3."'A2 A=AB*(A1+A3)*3./8. END IF IF(BETA.GT.(ZETA-1.5)) THEN AI=SA+I. A2=l .- 1 J(BETA-0.5)"3 162 A3=A2I3. A=AB‘(A1+A3)*3./8. END IF IF((BETA.GT .1 .5).AND.(BETA.LE.(ZETA-l .5))) THEN AI=7./3.-1./3./(BETA-0.5)""“3 A2=1.-1./SA"3 A3=A2/3. A=AB*(A1+A3)*3./8. END IF RETURN END C Use subroutine CUBIC as per flat wall program 163 C PORE -PURE GAS C *ltttfitttfitifitt***t********¥*t*¥**************#******************* C234567 Program to calculate surface excess or adsorption isotherms The program uses Pcng-Robinson equation of state, with a modified a to account for exclusion. The strength of the adsorption potential and its dependence on position is given by Psi. The fluid-fluid potential uff = u total - 11 ads. From this potential the various fluid properties are calculated. The variables used are defined below. All units used are SI unless otherwise stated. 0000000000000 DIMENSION H90(30000),H91(30000),H92(30000),H93(30000),H94(30000) DIMENSION H3(30000),H31(30000),H32(30000),H33(30000),H34(30000) DIMENSION AC(30000),RHO(30000),D15T(30000) REAL TC.PC,R,PLIM,AB.A,BB.B.T.OMEGA,DELP COMMON R.TC.PC.OMEGA.AC PSI(ETA)=0.382*3.14159’3.141S9'SIGFW'SIGFW’EPSFW*(3.‘SIGFW"*4“ I(H3(K)"(RFW/(RFW*RFW-ETA‘ETA'SIGMA*SIGMA))"4+H31(K)*((RFW+ALPHA) */((RFW -1-ALPHA)"l I*2-ETA *ETA *5lGMA‘SIGMA))**4+H32(K)*((RFW+ *2.‘ALPHA)/((RFW+2.*ALPHA)**2-ETA'ETA‘SIGMA'SIGMA))**4+H33(K) "((RFW+3.‘ALPHA)/((RFW+3.*ALPHA)**2-E'TA‘ETA *SIGMA*SIGMA))**4+ *H34(K)‘((RFW+4."'ALPHA)/((RFW+4.‘ALPHA)*‘2-ETA‘ETA‘SIGMA*SIGMA)) “‘4)-63./32."'SIGFW“10"(H90(K)"'(RFW/(RFW*RFW—ETA"'ETA*SIGMA*SIGMA)) "*10+H91(K)’((RFW+ALPHA)/((RFW+ALPHA)**2-ETA*ETA*SIGMA*SIGMA)) "' "" 10+H92(K)"((RFW +2. I"Al..PHA)/((RFW +2. l"ALPHA)"'"‘2-E’TA"'E1‘A"‘SIGMA"‘ 'SIGMA))*" 10+H93(K)*((RFW +3.‘ALPHA)/((RFW +3.*ALPHA)"2-ETA*ET A “SIGMA‘SIGMA))"10+H94(K)*((RFW+4.*ALPHA)/(CRFW+4.*ALPHA)**2- I"ETA"EI‘A"SIGMA"SIGMA))""" 10)) DO 10 1=1.sr.1 OPEN(UNIT=2,FILE='hyperS-9.dat',STATUS='UNKNOWN') READ(2,") ET,F0,Fl ,F2,F3,F4 C WRITE(*,*)ETAI J=ET*1000 H90(J)=F0 H9I(I)=Fl H92(I)=F2 H93(J)=F3 H94(I):F4 C WHIP—('a') J.Er.H90(J).H9l(J).H92(J).H93(J) 10 CONTINUE DO 15 M=1,8I,I OPEN(UNIT=4.FlLE='hypcr5-3.dat'.STATUS='UNKNOWN') READ(4,“) ET Al ,GO,G 1 ,GZ,G3.G4 J=ETA1"'1000 C 164 H31 (I)=G 1 H32(I)=G2 H33(I)=G3 H3(I)=GO H34(J)=G4 WRITE("‘."‘) J.ETAI .H3(J).H31(I).H32(J).H33(J) 15 CONTINUE 0000000000000000 0o~0 DO 6 MA=I.81.1 OPEN(UNIT=3,FILE='ethdia5.dat'.STATU5='UNKNOWN') READ(3."') ETAIANEW =ETA1 "' 1000 AC(J)=ANEW WRITE(‘.*) JETAIACU) CONTINUE write(*.*) AC(O) R = Gas constant I/K/mol Tc = Critical Temperature K Pc = Critical Pressure N/m2 OMEGA = Acentric Factor SIGMA = Sigma fluid-wall in m. T = Temperature K - FLIM = Bulk fugacity limit. EPSFW = Fluid-wall potential well depth in K SIGFW = Sigma fluid-wall in Angstroms ALPHA = Spacing of graphite planes / Sigfw SIGFF = Sigma fluid-fluid in Angstroms SIGWW = Sigma wall-wall in Angstroms AB = Bulk Peng-Robinson 'a' BB = Bulk Pong-Robinson 'b' ALPHA = Spacing of graphite planes / Sigfw SIGFF = Sigma fluid-fluid in Angstroms WRITE("'."')' Plim ,T,EPSILON' READ("',"') PLIM.T.EPSFW R=8.3 14 TC=282.4 PC=50.4E5 OMEGA=0.089 SIG MA=4.22 5 IGWW=3.4 ALPHA=3.35 RFW= 1 2.25 SIGFW=(SIGWW+S IGMA)/2. SIGGS=SIGMA*IE-10 Write parameters to files OPEN(UNTT=7,FTLE='denpm.dat'.STATUS='UNKNOWN') OPEN(UNIT=8,FILE='adsorp.dat'.STATUS='UNKNOWN') WRITE(7.") ' Ethylene at ',T.‘ K with an epsilon of ',EPSFW WRITE(8,"') ' Ethylene at ',T.‘ K with an epsilon of ',EPSFW 00.10 M 0000 165 WRITE(8,*)' RFW is ',RFW. ' angstroms '. 'R is RFW - 1.7' WRITE(8.*) WRI’I'E(8.")' Pbar',', ‘,'Excess um/m2'.'. ','Den molcs/cc' WRITE(8.") 0.', '. 0.'. '. 0 TR=T/T C FOMEG=O.37464+I .54226’OMEGA-0.26992*(O MEGA "2) Calculate a bulk and b AB=0.45724"(R*TC*(1+FOMEG*(I-SQRT(TR))))**2/PC BB=0.07780*R*TC/PC PCB=PC/1.0E5 IF(TR.GE. 1) GOTO 2 CALL FSAT(AB,BB.T.FUGS,PS,VV,VL) DELkPLIM/40. Loop for bulk fugacity PB=0.0E5 I=0 FOR EACH BULK FUGACITY VALUE PB=PB+DELP PB=PLIM B=BB A=AB First calculate bulk density (DENB) and bulk pressure (BP) CALL BVCAL(A,B,T,PB,PS,V,FB) CALL PV(A,B.T,V,P) DENB=1.0/V BP=P WRITE(7,"')'P = '. P.‘ bar' WRITE(7.‘)'DENB = ',DENB.‘ gmol/m3'.’ FB = ',FB.‘ bar' WRITE(7,") WRITE(7.')' ETA',', DEN gmol/m3'.'. ‘,'Log(Fugacity)' Start iterating in ETA DELETA=0.025 E‘I‘A=2.025 EXCESS=0.0 AMOUNT=0.0 AMOUNTI=0.0 KDEL=25 K=2025 L=0 ETA=ETA~DELETA K=K-KDEL Calculate local fugacity Use the following formula to get the local fugacity (F). for a given position ETA. F=FB"'EXP(PSI(ETA)/I') 0000 C 166 ALNFB=ALOG(FB) ALNF=ALNFB+PSI(ETA)/T PSIV=PSI(ETA)/T Calculate local a WRITE(".*)K,H3(K).H3I(K).H32(K).H33(K) WRITE(*,*)K.H90(K).H91(K),H92(K).H93(K).PS1(ETA) WRITE(*.*) K. AC(K) AC(O)=0.9262953 CALL ACALC(ETA,K,AB.A) IF(TR.GE.1.) GO TO 20 CALL FSAT(A.B.T.FUGS.PS,VV,VL) Using local parameters calculate V and local density DENL 20 CALL VCALC(A.B.T.ALNF.BP.PS.V) C C C C C DENL=1 .0/V Using Rectangular Rule: . EXCES S=SIGGS *(DENL-DENB)"DELETA* l .E6+EXCESS Amount is den*r*dr*2 Pi L; 2 Pi L = 988/Pore Radius. Amt.=denl*siggs'eta*siggs‘deleta‘988/(2.5'siggs) [mmol/g] AMOUNT =SIGGS*DENL'ETA‘DELE’TA"988/2.5* l .e3+AMOUNT RHO(L)=DENL DIST(L)=ETA DENLG is the density in gmoles/cc DENLGzDENUI .0E6 .WRITE (7.*)ETA.’.'.DENL.'.‘.ALNF L=L+l IF(L.LT.81) GOTO 5 CALL SIMP(RHO.DIST.EXI.AMT1.L,DENB.SIGGS.A3) BP=P/ I .0E5 DENBG=DENB/1.E6 WRITE(8."') BP, ',', EXI . '.'.AMTI .'.',DENBG WRITE(6.*) BP, ',', EXCESS. ',',AMOUNT,'.',DENBG WRITE(6.*) BP, ',',EX1.',‘. AMT] .A3 J=I+l IF (J.LT.40) GOTO 3 102 FORMAT(1X.F8.2.2X.F12.5.2X.F10.5) END CIit.Ililtlttttlttttittfitttttttlfitfilttt*ttIII!!!****************** C234567 C SUBROUTINES Ct...fittitl##1##tit.tit.*tfitttlttttitltttifi*tt*tfitttit#********** C THIS IS A SIMPSON'S RULE SUBROUTINE 167 SUBROUTINE SIMP(RHO.DIST,EX1.AMT1.L.DENB,SIGGS.A3) DIMENSION RHO(30000),DIST(30000) E1=0 A1=0 E2=0 A2=0 DO 250 I=0.L-1 A3=A3+SIGGS I"RHO(I)"'DIST(I)“0.025 I"988./2.5"' 1 .E3 E3=E3+SIGGS ‘(RHO(I)-DENB)*0.025*1 .E6 250 CONTINUE DO 200 I=1,L-2,2 E1=E1+(RHO(I)-DEN B) A 1=A1+RHO(I)*(DIST(I)"SIGGS) 200 CONTINUE DO 225 1=2.L-3.2 E2=EZ+(RHO(1)-DENB) A2=A2+RHO(I)‘(DIST(I)’SIGGS) 225 CONTINUE 000000 0000000 EXI=RHO(O)-DENB+RHO(L-1)—DENB+4.*EI+2.*EZ EXI=0.025*SIGGSI3.*EXI'I.E6 AMTI=0.025/3.*(DIST(L.I)*SIGGS ‘RHO(L-l)+4.*Al+2.*A2) AMT I=988./2.5"AMT1 *1.E3 RETURN END Use subroutine VCALC as per fiat wall program Use subroutine DPDV as per fiat wall program Use subroutine VICALC as per flat wall program Use subroutine V3CALC as per fiat wall program Use subroutine PV as per fiat wall program Use subroutine ARANGE as per fiat wall program SUBROUTINE ACALCCETA.K .AB.A) This subroutine calculates the van der Waals a term after taking into account the effect of exclusion. AB = Value of a in the bulk ETA = reduced distance from the center of wall The main program sends a reduced distance ETA from the center of the wall molecule, which is the basis for the integrated 9-3 potential. DIMENSION AC(30000) 168 COMMON R.TC.PC.OMEGA.AC A=AB‘AC(K) IF (AC(K).EQ.0.0) THEN K 1=INT(K/lOO) K2=KI '100 AK2=FLOAT(K/100.-KI) K3=(K 1+ I)"'100 , AC(K) = AC(K2)+(AC(K3)-AC(K2))*AK2 WRITE(*.*) K.K1.K2.AK2.K3.AC(K) A=AB’AC(K) ENDIF RETURN END C Use subroutine CUBIC as per fiat wall program 000000000000 0 g. 169 PROGRAM FOR BINARY MIXTURES on SLITS Define Psi as a statement function Use a 10-4 potential where PSI is the negative of the intermolecular potential. SIGGSI and SIGGSZ are sigma fluid-wall for components 1 & 2. in Angstroms. EPSGSI & 2 are the fluid-wall potential in K EXCI & 2 are the excess in micro mol/m"2 AMTI & 2 are the amount adsorbed in mmol/g YI & 2 are the compositions of the two components in the adsorbed phase PI & 2 are the fugacity coefficients for the two components Fl & 2 are the fugacities p is the pressure in Mpa TC, PC and W represent the critical temperature. pressure and omega’s This program is to calculate the adsorption in a mixture REAL K12 CHARACTER RFILE*80 PSI1(ETA)=4.0*3.1415926*O.382*SIGGSI*SIGGS1‘EPSGSI*(-O.2 1/ETA" l O-l-O.5/E'I‘A"'"‘4+0.5/(E'I'A+ALPHA)"'"'4+-0.5/(ETA-l-2.0"l ‘ALPHA)"4+O.5/(ETA+3.0*ALPHA)"4+O.5/(ETA+4.0‘ALPHA) Oit4) P512(ETA)=4.0*3.1415926‘0.382*51G652‘SIGGS2*EP5GS2*(-O.2 l/ETA“10+0.5/ETA"*4+0.5/(ETA+ALPHA)‘*4+0.5/(ETA+2.0"' *ALPHA)"4+O.5/(ETA+3.0’ALPHA)‘*4+0.5/(ETA+4.0*ALPHA) tfit4) PSI3(XI)=4.0*3. 1415926‘0.382*51GG5 1 "SIGGSI I"EPS GS 1 I"(~02 1/X1**l 0-t~0.5/XI"'"‘4¢l-t'0.S/(XI+ALPHA)"“"4+0.5/(XI+2.O"l "ALPHA)"4+0.5/(XI+3.0*ALPHA)“4+0.5/(X1+4.0‘ALPHA) $.14) PSI4(XI)=4.0*3. 14 l 5926*0.382*SIGGSZ‘SIGGSZ*EPSG52*(-O.2 l/XI”10+0.5/XI"4+0.5/(XI+ALPHA)"4+0.5/(Xl+2.0" ‘ALPHA)"4+0.5/(Xl+3.0*ALPHA)**4+O.5/(Xl+4.0*ALPHA) OOI4) PSIA(ET A)=PSI 1 (ET A)+PSI3(XI) PSIBfETA)=P512(ETA)+PSI4(XI) SIGGS 1=4.0 SIGGSZ=4.0 ALPHA:3.35/SIGGS l R = 8.314 TC2=282.4 PC2=5 .04 W2=0.089 170 TC1=364.9 PC1=5.49 W1=0.13 K12=0 PC2 = 1000000‘PC2 PCI = IOOOOOO‘PCI 2 PRINT“. 'ENTER T(K). P(MPA),EPSI, EPSZ, WIDTH' READ“. T.P.EPSGS I .EPSGSZZETA P=1000000*P 3 PRINT“, 'ENTER Yl' READ", Y1 EXC1=0.0 EXC2=0.0 AMT1=0.0 AMT2=0.0 TYPE'(X.A.$)'.'ENTER THE FILE NAME FOR RES ULTS :' ACCEPT'(A)'.RFILE OPEN(UN1T=2.NA ME=RFILE.TYPE='UNKNOWN') Y2=1~Y1 YIIN=YI Y21N=Y2 DDYI = 1E-5 DDY2 = IE-S TR1=T/TC1 TR2=TITC2 " SET DELTA P FOR PARTIAL DERIVATIVES C DDP=.01‘P " CALCULATE COMPRESSIBILITY FACTOR (Z) “ CALCULATE A1.A2. THEN A CALL ACALC(R.WI.TC1.TR1.PC1.A11) CALL ACALC(R.W2.TC2.TR2.PC2,A22) CALL BCALC(BI,TC1.PC1 .R) CALL BCALC(B2,TC2.PC2.R) A12 = SQRT(A1 1‘A22)’( I-K12) CALL AMIX(YIIN.Y21N,A11.A22,A12.A) CALL ASTAR(A1 1.A22.A12.A.ASI.AS2.A512.AS.P.R,T) CALL BMIX(Y 11N.Y21N.BI.B2.B) CALL BSTAR(B 1 .BZ,B.BS I ,BSZ.BS,P,R.T) CALL SCUB1C(AS.BS.R1.R2.R3.IFLAG) IF(IFLAG.EQ. 1) z=R1 IF(IFLAGEQ.2) THEN z.-.-Rt IF(R2.GT.RI) Z=R2 END IF IF(IFLAG.EQ.3) THEN 171 CALL ARANGECRI ,RZ.R3) Z=Rl END IF CALL PH(YIIN,Y21N.A51.A512.A52.A5.BSI.BSZ.BS.Z.P1.P2) C WRITE("',*) 'A & B star ',AS,BS.'FUG’.P1,P2 DENB=PfZlRfT PBULK=PI 1000000 WRITE(2.*) 'T ='.T.' K'.'P ='. PBULK.‘ MPa' WRITE(2."') 'BULK DEN ='. DENB,‘ gmol/m"3'.' Width = ',ZETA WRIT'E(2,"')'Epsilon 1 & 2 ='. EPSGSI .EPSGSZ WRITE(2,*) 'inital mol. frac. Y1 = ',Yl ,' Y2 = '. Y2 FIBULK=YIIN*P1*P F28ULK=Y21N*P2*P PETA=P C ------------- Start iterating in ETA BET A=ZETA/2.+(ZET A/2.-O.5)/ 100. DELETA=(ZETA/2.-0.5)/100. L20 WRITE (2,102) 5 BETA=BETA-DELETA ET A=(BETA‘2J'SIGFF+SIGWW)/(SIGFF+SIGWW) XI=(ZETA*SIGFF+SIGWW/2.-BETA"SIGFF)/5IGGS1 "' SET DELTA P FOR PARTIAL DERIVATIVES DDP=.01*PETA GOLD=IE5 K=0 H = 0 CALL AZCALC(BETA.A1 1.A1.ZETA) CALL AZCALC(BETA,A22.A2.ZETA) A12 = SQRT(A1 l"A2)"'(1-K 12) F1 = FIBULK‘EXPO’SIAG'TTAVT) F2 = FZBULK*EXP(PSIB(ETA)/T) P=Fl/P1+F2/P2 C WRITE("',") 'F1.F2.P'.F1.F2,P 10 CALL AMIX(Y1.Y2.A1,A2.A12,A) C WRITE(",") 'A='. A CALL ASTAR(A1.A2.A12.A.ASI.A52.A512.A5.P.R,T) CALL BM1X(Y 1.Y2.BI.BZ.B) C WRITE("'.*) 'B=', B 172 CALL BSTAR(B I ,BZ.B,BS l .BSZ.BS.P.R.T) CALL SCUBIC(AS.BS.RI ,R2,R3.IFLAG) IF(IFLAG.EQ.1) THEN ZI=RI CALL PH(Y1,Y2.ASI.A512,A52.AS.BS 1.852.851] ,P1,P2) CALCULATE OBJECTIVE FUNCTIONS G1 = Y1*P1"‘P/Fl-1 02 = Y2'P2’P/F2-l G3 = l - Y1 - Y2 G0=SQRT(GI"'GI+ G2*G2 + G3*G3) NROOT=1 GOTO 25 ENDIF IF(IFLAG.EQ.2) THEN 2] =R 1 Z2=R2 IF(R2.GT.R 1) THEN ZI =R2 22=R 1 ENDIF EN DIF IF(IFI..AG.EQ.3) THEN CALL ARANGE(R1.R2.R3) Zl =R 1 22=R3 EN DIF CALL PH(Y1.Y2,ASI.A512.A52.AS.BS1352.35.21.PIA,P2A) CALCULATE OBJECTIVE FUNCTIONS GIA = Y1*P1A*P/Fl-l G2A = Y2*P2A*P/F2-1 G3A = l - Y1 - Y2 GOA=SQRT(GIA*GI A+ GZA*G2A+ G3A‘G3A) WRITE("'.")'GIA.G2A'.G1A.GZA GO=GOA GI=GIA GZ=G2A G3=GBA 2:21 P1=P1A P2=P2A NROOT =1 IF(ZZLTBS) THEN PRINT". 'SKIPPING 22' GOTO 25 EN DIF CALL PH(Y 1.Y2,A51.AS12,A52.AS.BSI.BS2.BS,ZZ,PIB,PZB) CALCULATE OBJECTIVE FUNCTIONS 618 = Y1‘P18‘P/F1-1 173 623 = Y2’P28‘P/F2-1 038 = 1 - Y1 - Y2 GOB=SQRT(GIB*GIB + G2B‘G2B + G3B*G3B) IF (GOA.GT.GOB) THEN GO=GOB G 1:0 1 B GZ=G2B G3=G3B 2:22 P I =P I B P2=PZB I NROOT=2 ENDIF 25 CONTINUE PRINT*,'GO=',GOA,GOB ,'NROOT='.NROOT,'IFG='.IFLAG F1CALC=Y1*P1"‘P F2CALC=Y2"‘P2"P YT=Y I +Y2 WRITE(","') F1CALC.F1,P WRITE(*,*) F2CALC,F2,H PRINT", 'ETA,Y I ',ETA,Y2 IF (K.LT.5.AND.GOLD.LT.GO) THEN K=K+I Y I=Y I -DYI Y2=Y2~DY2 P=P-DP DY] = DY 1/2 DY2 = DY2/2 DP = D?” GOTO 100 END IF K=0 GOLD=GO 9999 FORMAT(4GI$.5.IX,I3) 000 C IF(GO.LT.2E-2) THEN lF(GO.LT.1E-3) THEN C WRITE (*,"‘) 'Bubblc Pressure Calculation' C WRITE (‘,101) C WRITE (',103) X1.X2,T DEN=P/Z/8.3 14/1‘ PETA=P EXC1=SIGGS 1 *(Y I *DEN-Y I IN‘DENB)*DELE'TA* 1 .E-4+EXCI AMTI=Y1*DEN‘SIGGSI‘DEETA‘988.*1.E-7+AM'T1 EXC2=S 1668 1 *(Y2‘DEN-Y2IN‘DENB)*DELETA *1 .E-4+EXC2 AMT2=Y2"DEN"S 1068 1 ‘DELE’TA‘988J‘I .E-7-I-AM'T2 WRITE (‘,104) Y1,DEN,P,BETA,H WRITE (2,104) Y1 ,DEN,P,BETA,H 101 FORMAT( I 0X,'X I',lOX,'X2',10X,'T—K') 174 102 FORMAT(4X,'Y I ',7X,'DEN'.9X,'PRESS3,9X,'ETA'.8X,'# ITER') 104 FORMAT(F8.5,2X,ZG12.5.2X.GI2.5.2X,F5.l) GOTO 200 ENDIF YY1= Y1 + DDYI YY2 = Y2 + DDY2 PP=P+ DDP "' CALCULATE DERIVATIVES WRT P CALL ASTAR(A1,A2,A12,A,ASI ,ASZAS 12,AS,PP,R,T) CALL BSTAR(B 1 ,BZ,B,BS 1.BSZ,BS ,PP,R,'I') CALL SCUBIC(AS ,BS,R 1 ,R2,R3,IFLAG) IF(IFLAG.EQ.1) THEN Z=R1 CALL PH(Y 1,Y2.ASI ,ASI2,A52,AS.BSI ,B82,BS .Z,PN1,PN2) GOTO 30 ENDIF IF(IFLAG.EQ.2) THEN B IG=R1 S MALL=R2 IF(RI .LT.R2) THEN BIG=R2 S MALL=R1 ENDIF R 1 :8 IO R3=S M ALL EN DIF IF(IFI..AG.EQ.3) THEN CALL ARANGE('RI ,RZ,R3) ENDIF IF(NROOTEQ. 1) THEN Z=R 1 CALL PH(Y 1,Y2,ASI ,AS 12,ASZ,AS,BS I ,BSZ,BS ,Z,PN 1,PN2) ELSE Z=R3 CALL PH(Y 1,Y2,ASI,ASI2,ASZ,AS,BSI,BSZ.BSZ.PN1.PN2) ENDIF 30 D! DP = (Y1*PN1*PP-Y1‘Pl"P)/DDP/Fl D2DP = (Y 2*PN2*PP-Y2*P2*P)/DDP/F2 DBDP = 0 " CALCULATE DERIVATIVES WRT Y1 CALL AMIX(YY1,Y2,AI,A2,A12,A) CALL BMIX(Y Y1 ,Y2,B I ,BZ,BV) CALL AS'TAR(AI,A2,A12,A.AS1.AS2.ASI2,AS,P,R,T) CALL BSTAR(B 1 ,B2,B.BS 1 ,BS2.BS,P,R,T) CALL SCUBIC(AS ,BS,RI ,R2,R3,IFLAG) 175 IF(IFLAG.EQ.1) THEN Z=R1 CALL PH(YYI .Y2.As 1 .As 12,A82,AS,BS 1 352.351.9111 .PN2) 0010 40 ENDIF IF(IFLAG.EQ.2) THEN BIG=R1 SMALL=R2 IF(RI.L'T.R2) THEN BIG=R2 SMALL=R1 ENDIF RI=BIG R3=SMALL EN DIF IF(IFLAGEQ.3) THEN CALL ARANGE(R1.R2.R3) EN DIF IF(NROOTEQ. 1) THEN Z=Rl CALL 911(le ,Y2,AS 1.As 12,A32,As,as1,Bsz,Bs,zJ>N1.1>N2) ELSE Z=R3 CALL PH(YYI ,Y2.AS 1 ,As 12,A52.As 351,352.35 .Z,PN1,PN2) ENDIF 40 D1 D1 = (PNI l"YYI *P-PI‘YI *P)/DDYI/F1 D201 = (PN2'Y2‘P-P2‘Y2‘PVDDY IIF'Z D3D1 = -1 "' CALCULATE DERIVATIVES WRT Y2 CALL AMIX(Y 1,YY2,A1,A2.A12,A) CALL BMIX(Y1,YY2,BI,BZ,B) CALL ASTAR(AI.A2.A12.A.AS1.A52,ASI2,AS.P,R.T) CALL BSTAR(B 1 ,BZ,B.BS I ,BSZ,BS,P.R.T) CALL SCUBIC(AS ,BS,R 1 ,RZ,R3,IFLAG) IF(IF'LAGEQJ) THEN Z=R1 CALL PH(Y I ,YY2.AS 1,AS 12.A82,AS,BSI ,BSZ,BS,Z,PN1.PN2) GOTO 50 ENDIF IF(IFLAG.EQ.2) THEN BIG=R I SMALLzRZ IF(R1.LT.R2) THEN BIG=R2 SMALL=RI ENDIF R 1=BIG 176 R3=SMALL ENDIF IF(IFLAGEQ.3) THEN CALL ARANGE(R1,RZ,R3) EN DIF IF(NROOTEQ. 1) THEN Z=R1 CALL PH(Y 1,YY2,AS 1 ,AS l2.A52.AS ,BS 1 ,BSZ,BS Z.PN1,PN2) ELSE Z=R3 CALL PH(Y 1 .YY2,ASI ,A812.A82.AS ,BS 1 ,B82,BS .Z,PN1.PN2) ENDIF 50 D1D2 = (PNI *YI l"P-PI I"Y1 'P)/DDY2/FI D2D2 = (PN2‘YY2‘P-P2’Y2'PVDDY2/F2 D3D2 = -1 * CALCULATE INCREMENTS CALL INV(D 1 DP,DZDP.D3DP,DI D1 ,DZDI,D3D1,D1D2,D2D2.D3D2,DET) DY1 = -GI"DID1 - G2‘D1D2 - G3‘D1DP DY2 = -Gl"D2D1 - GZ‘D2D2 - G3‘D2DP DP = -Gl"‘D3D1 - 62*D3D2 - G3‘DBDP PRINT‘, ETA,Y1,DY1,P,DP IF(ABS(DY1).GT.YI) THEN PRIN'T*,'WARNING, LARGE DY',DYI DYI=0.5*Y1*DY1/ABS(DY1) DY2=-DY1 C P=9.E7 C DP=0 END IF 100 Y1 = Y1 + DY1 Y2 = Y2 + DY2 P = P «1» DP IF(Y1.LT.0)Y1=0 IF(Y1.GT.1)Y1=1. IF(Y2.LT.0)Y2=0 IF(Y2.GT. 1)Y2= I . H=H+1 IF (H.LT.200) GO TO 10 PRINT", 'ITERATIONS EXCEEDED' C DDYI=1E-3 C DDY2=1E-3 C DDP=.01"P C GOTO 10 200 CONTINUE L==L+l IF(L.LT.101) GOTO 5 WRITE(",") EXC1,EXC2,DENB,AMT1,AMT2 177 WRITE(2,"') EXC1,',',EXC2,',',DENB WRITE(2,"') AMT 1.',',AMT2 PR1NT*,'ENTER 1 FOR SAME T, DIFFERENT x1' PRINT", 'ENTER 2 FOR NEW T AND NEW x1' PRINT‘, ENTER 0 TO QUIT READ", IDB IF (IDB.EQ.1) GO To 3 IF (IDBEQ.2) GO To 2 END .**ilfitfifitttfififitltil*Ititltlifi*ltt***********************¥************* " SUBROUTINES BELOW 1!!! * SUBROUTINE ACALC(R,W,TC,TR,PC.AN) FW = 0.37464 + 1.54226‘W - .26992 " W**2 AN = 0.45724'R“2"TC"2"(1+FW*(1-SQRT(TR)))"2/PC RETURN END SUBROUTINE AMIX(MI ,M2.A1.A2.A12,AA) REAL AA, A1, A2, A12, M1, M2 AA = M1"2"A1 + 2*M1‘M2‘A12 + M2"2 * A2 RETURN END SUBROUTINE ASTAR(AI ,A2.A12,A,AS 1 ,AS2.AS l2.AS .P,R,T) AS] = AI‘P/(R"2*T”2) A52 = A2'P/(R"2*T”2) A812 == A12’P/(R"2*T“2) AS = A‘P/(R"2"T"2) RETURN END SUBROUTINE BCALC(B,TC.PC,R) B = 0.07780‘R'TC/PC RETURN END SUBROUTINE BMIX(M1,M2,B 1 ,B2,BB) REAL BB, Bl. B2, M1, M2 BB = M1’BI+ M2*B2 RETURN END SUBROUTINE BSTAR(BI,132,B.Bs1 ,BSZ,BS,P,R,T) 1351 = Bl‘P/(R‘T) 1332 = BZ‘P/(R‘T) as = 3*9/(Rm RETURN END SUBROUTINE TM2(BS,T2) T2 = BS - 1 WRITE (‘3') T2 178 RETURN END SUBROUTINE TM1(AS,BS,T1) T1 = AS - BS"(2+3"BS) WRITE (‘,‘) T1 RETURN END SUBROUTINE TMO(AS,BS,T0) '11) = BS“(BS"'*2 + BS - AS) WRITE(*,“‘) TO RETURN END SUBROUTINE PH(MI ,M2,AS I ,A812A82,AS,BS 1 ,Bsz,Bs,z,I>1,I>2) REAL M1,M2 Q = M1’ASI+ M2*A512 CALL PHICLC(BS l ,AS.BS.Z.Q.P1) Q = Ml‘ASlZ + M2-Asz CALL PHICLCCBSZ,AS,BS.Z.Q.P2) RETURN END SUBROUTINE PHICLC(B,AS,BS,Z.Q.PHI) TERMI = (BlBS)*(Z-1) TERMIA = (2*Q/AS) - (BIBS) TERMIB = 2"Q/AS TERM2 = (2*Q/AS - B/BS)"'(AS/(2*SQRT(2.0)"BS)) TERM3 = I/(Z-BS) TERM4 = ((Z-(SQRT(2.0)-1)*BS)/(Z+(SQRT(2.0)+ 1 )"‘BS )) WRTTE(*.") BS, TERMIA, TERMIB WRITE(*,") TERM2, TERM3, TERM4 PHI = TERM3*(TERM4“TERM2) I"EXP('I'ERM I) RETURN END SUBROUTINE SCUBIC(ASBS.RI.RZ.R3.IFLAG) CALL TM2(BS,T2) CALL TM1(AS,BS,T1) CALL TMO(AS,BS,'m) WRITE (*,"') T2,TI ,TO CALL CUBICCI‘2,TI,T0,RI,R2.R3,C1,C2,C3,IF'LAG) WRITE(",*) IFLAG WRITE("',") 'R1 R2 R3' WRITE("',") R1,R2,R3 RETURN END 179 C SUBROUTINE ARANGE(R1,RZ,R3) C PROGRAM TO PUT 3 NUMBERS IN DESCENDING ORDER DO 20 1:13 ’ IF(R2.GT.R 1) THEN TEMP=R I R I =R2 R2=TEMP ENDIF IF(R3.GT.R2) THEN TEMkR2 R2=R3 R3=TEMP ENDIF 20 CONTINUE RETURN END C SUBROUTINE AZCALC(BETA.AB,A.ZET A) SA=ZETA.0.5-BETA IF (BETA.LE.I.5) THEN A 1=BETA+0.5 A2=I.-I./SA*"'3 A3=I./3."'A2 A=AB *(AI+A3)"3./8. END IF IF(BETA.GT.(ZETA-1.5)) THEN AI =SA+1. A2=1.-1J(BETA-0.5)"3 A3=A2/3. A=AB“(A1+A3)*318. END IF IF((BETA.GT .1 .5).AND.(BETA.LE.(ZETA-I .5))) THEN A1=7./3.-1 l3./(BETA—0.5)"3 A2=1.-1./SA“'*3 A3=A2/3. A=AB*(A1+A3)*3./8. END IF RETURN END C 180 SUBROUTINE INV(DIDP,D2DP,D3DP,D1D1,D2D1,D3D1,DID2,D2D2,D3D2,DET) C011 = D2D2’D3DP-D3D2‘D2DP C012 = -(D2D1*D3DP-D2DP*D3DI) C013 = D2D1*D3DZ-D3D1"D2D2 C021 = -(D1D2*D3DP-D3D2*D1DP) C022 = D1DI*D3DP-D3D1"'D1DP C023 = -(D1 DI *D3D2-D3DI *D1 D2) C031 = D1D2’D2DP-D2D2’DI DP C032 = -(D1D1*D2DP-D2D1 I"DI DP) C033 = D1D1’D2D2-D2D1'D1D2 DET: D1DI'C011+ D1D2‘C012 + DIDP‘COI3 WRITE("',*) 'DET' WRITE("',*) DET DIDI = C01 1/DET DID2 = C021/DET D1DP= CO31/DET DZDI = C012/DET D2D2 = C022/DET D2DP = CO32/DET D3Dl = COI3/DET D3D2 = COB/DET D3DP = C033/DET RETURN END Use Subroutinc CUBIC as per other C C C 15 17 181 ACALIDOC This is a program to calculate the value of the vanderwaals 'a' in a cylindrical pore for T] varies from R-3 sigma/2 to R-sigma/‘Z IMPLICIT DOUBLE PRECISION(A-Z) INTEGER I,J,I( WRITE(*,"‘) 'ETAI',' SIGMA'.' PORE SIZE IN NUMBER OF SIGMA' READ("'.*) ET A1,SIGMA.P R1=SIGMA *ETAI =SIGMA'P WRITE(*,*) R,Rl CASE 1 rl varies from R-3 sigma/2 to R-Sigma/Z A 1 =0.0 P1=3.1415926535897932 z1=(SIOMA*SIOMA-cR-SIGMA/2.-R1)"2) WRITE(*,") 21 ZI=DSQRT(ZI) WRITE(*,*) 21 DO 10 1:0.1000 bl'Zl/IOOO. WRITE(*,*) z B=Rl *RI+SIOMA*SIGMA-z*z B I=(R-SIGMA/2.)*(R-SIGMA/2.) 32=SIGMA*SIGMA.z*z WRITE(*,*) 132 IF(BZ.EQ.0.0) THEN WRITE(*,*) '32 EQ 0' GO TO 15 ENDIF B3=DSQRT(B2) A=(B-B I )flJRI/B3 WRITE(*,"‘) A IF(A.GT.1.0) THEN WRI'I'E(","') 'A.GT.1.0 ',Z,A A=1 .0 ENDIF IF(A.LE.-1.0) THEN WRITE(*,*) Z,A =—I .0 ENDIF F=DACOS(A) GO TO 17 =PI A1=F’DSQRT(SIGMA*SIGMA-(R-SIGMA/2.-R1)**2)/1000.+A1 10 CONTINUE A1=Al/4./(SIGMA**4) 182 WRITE(*,"') 'A1= ',AI DO 20 J=0.10(X) ZI =DSQRT(S IGMA‘SIGMA-(R-SIG MHZ-R I )**2) Z=J*ZI/l 000. D0 30 K=0,1000 B=RI *R 1+SIGMA"'S IGMA-Z‘Z B 1=(R-SIGMA/2.)*(R-SIGMA/2.) BkSIGMA‘SIGMA-l‘z IF(BZ.EQ.0.0) GO TO 25 B3=DSQRT(BZ) A=(B-B ”fl/R 1/33 IF(A.GT.I .0) THEN WRITE(*,"') 'AA GT 1.0 'Z.A A=I .0 ENDIF IF(A.LT.-I.0) THEN WRITE(*,*) Z, A =-1.0 ENDIF GO TO 27 25 =-I.0 27 THETA=DACOS (A)*K/I (XX). C WRITE(*.*) Z, A,THETA C=Rl *DCOSCT HET A) C1=(R-SIGMAI2.)"(R-SIGMA/2.) C2=R1 I"R1"DSIN('THE’I‘A)"‘DSIN('T1-IE'I‘A) C3=DSQRT(C 1 -C2) C4=C+C3 F I =C4 *C4+Z"Z F=I JF III: I AA=F"'ZI/1000."DACOS(A)/I000.+AA 30 CONTINUE 20 CONTINUE AA=AA/4. WRITE(*,*) 'AA= ',AA =DSQRT(SIGMA*SIGMA-(R-SIGMA/Z-R1)"'*2) DO SO K=0.IOOO THETA=P1"K/1000. C=R1 ‘DCOS (THET A) C 1=(R-SIGMAf2.)"(R-SIGMA/2.) C2=Rl"Rl"DSIN('1'I-IEI‘A)*DSIN(TI-1ETA) C3=DSQRT(CI -C2) C4=C+C3 C5=C4*C4 F1=SIGMA/2JC5/(C5+SIGMA"SIGMA) F2=Yl2JC5/(C5+Y"Y) 183 F3=DATAN(SIGMA/C4)/2./CS/C4 F4=DATAN(Y/C4)/2./C5/C4 F=F1-F2+F3-F4 AB=F*PI/1000.+AB 50 CONTINUE AB=PI*(SIGMA-Y)/4./SIGMA**4-AB/4. WRITE(*,‘) 'AB = ',AB C PART 21.e. 2 going from sigma to infinity C D0401=0,1000 THETA=I"'Pl/1000. B=R l *DCOSCI‘HETA) B 1=(R-SIGMA/2.)*(R-SIGMA/2.) Bth’Rl *DSINCI'HETA)’DSIN(THETA) 83=DSQRT(Bl-82) A=B+B3 F1=PI/4./NA/A F2=SIGMNZJAIAKA *A+SIGMA’SIGMA) F3=DATAN(SIGMA/A)/2JAIA/A =Fl-F2-F3 WRITE(*,*) Fl ,F2.F3 A2=F‘PI/1000.+A2 40 CONTINUE A2=A2/4. A3=PI/1ZJSIGMA/SIGMNSIGMA-A2 WRITE(*,*) A2,A3 A4=A1-AA+AB+A3 A5=A4*3.*SIGMA‘SIGMA‘SIGMA/PI WRITE("'."‘) 'A/ABULK= '. A5.A4 END _ ‘ C C C C C C 184 ACALZJ: This is a program to calculate the value of the vandcrwaals 'a’ in a cylindrical pore rl varies from 0 to R-3sigma/2 IMPLICIT DOUBLE PRECISION(A-Z) INTEGER I WRITE(*,*) 'ETAI'.’ SIGMA'.’ PORE SIZE IN NUMBER OF SIGMA' READ(*,*) ET AI,SIGMA.P R1=SIGMA"'ETA1 R=SIGMA‘P WRITE(*,*) R,RI CASE 2 rl varies from 0 to R-3sigma/Z A1=0.0 PI=3.1415926535897932 D0 10 1:0.1000 THE'TA=I"'PI/1000. B=R1 *DCOSCTHET A) B 1=(R-S IGMA/Z.)‘(R-SIGMA/2.) B2=R 1 *RI *DSIN(T HETA)‘DSIN(THET A) B3=DSQRT(B 1-B2) A=B+B3 F1=A*A"A F=1./FI WRTTE("',") THETA .A,F F3=2.*A *A‘A F=Fl +F2/F3 . A1 =F‘PI/1000.+A I 10 CONTINUE A2=1./3.-SIGMA*SIGMA*SIGMA/16.*A1 A3=A2*3. WRITE("‘,"') 'A/ABULK ', A3.A1,A2 END C C C 185 ACAL3.F This is a program to calculate the value of the vandcrwaals 'a' in a cylindrical pore r1 = R-sigma/Z REAL ET A1 ,SIGMA,P.R,R1 .A,A 1 .A2 WRTTE(*."') 'E'TAI'. ' SIGMA',’ PORE SIZE IN NUMBER OF SIGMA' READ("',*) ETA1,SIGMA,P R1=SIGMA"E'TA1 =SIGMA‘P WRITE(*."‘) R,Rl CASE 3 T] = R-sigma/Z A1=0.0 PI=3. 1415927 DO 101:0.100 Z=I"'SlGMA/IOO. B=SIGMA‘SIGMA-Z‘Z Bl=SQRT(B) THETA=ACOS(Bl/2./Rl) F=TI-1ETA/4./SIGMA"4 WRI'I'E(*.‘) Z.A.F A1=F"SIGMAI100.+A1 10 CONTINUE WRITE("',"‘) 'A1= ',Al DO 20 1:1,100 Z=J*SIGMA/100. B=SlGMA’SlGMA-Z‘Z Bl=SQRT(B) THETA=ACOS(Bl/2./Rl) C=2."'RI*R1+Z"'Z Cl=4."Rl*Rl+Z"‘Z C2=CI M1 .5 C3=ClC2flflfl write(*,"') C,CI ,C2,C3 IF(Z.EQ.SIGMA) GO TO 25 C4=Z*TAN(THETA)/SQRT(CI) C5=ATAN(C4) GO TO 27 C5=PI/2. C6=C3"'C5 write(*,*) C3.C5.C6 D1=R1*R1*SIN(2‘THETA) D2=Z"'Z*CI*(2.*R1*RI+Z*Z+2.*RI I"R1 *COS(2."‘THETA)) I86 D3=Dl/D2 C WRITE("',"') D1,D2,D3 F=C6~D3 A2=F‘SIGMA/100.+A2 C WRITE(*,"') Z, A,THET A 20 CONTINUE A2=A2/4. AA=A 1-A2 WRTTE("',"') 'A2= ',A2,‘ AA: ',AA C PART 2 i.e. 2 going from sigma to infinity DO 40 1:0,2000 Z=SIGMA+I*SIGMA/ 100. B=R1*RI+Z*Z BI=4.“R1*R1+Z*Z 82:81 “1.5 FEB/82m A3=F*SIGMA/100.+A3 40 CONTINUE A3=A3*PI/8. A4=PII24JS TOMA/S IGMA/SIGMA A5=A4-A3 WRITE(*,"‘) A3,A4,A5 A6=AA+A5 A=A6*SIGMA‘SIGMA‘SIGMA'BJPI WRITE(*,*) 'A/ABULK: ', A6.A END Appendix 3 - Properties of Gases 'Table: Appendix 3 -1 Supercritical fihylene 40 F P(Psia) 500 600 700 800 900 1000 1 100 1200 1300 1400 1500 V(f13llb) 0.26035 0.18181 0.04754 0.04511 0.04368 0.04267 0.04187 0.04122 0.04067 0.04019 0.03977 P(MPa) 3.447367 4.1 3684 4. 82631 3 5.51 5787 6.20526 6.894733 7. 584207 8.27368 8. 9631 53 9.652627 1 0.3421 V(cm3lmol V (PR) 455.9669 318.415 83.2597 79.0039 76.49945 74.73058 73.32949 72.191 1 71 .2785 70.3872 69.65163 443.6109J 307.7956 93.08558 85.36646 81 .0538 78.03798I 75.72067 73.84304 72.26802 70.91416 69.72904 160 F P(Psia) 500 600 700 800 900 1 000 1 100 1200 1 300 1400 1 500 V(ft3llb) 0.29183 0.21835 0.16006 0.10129 0.05804 0.05119 0.04821 0.04637 0.04506 0.04404 0.04322 P(MPa) 3.447367 4. 1 3684 4.82631 3 5. 51 5787 6.20526 6.894733 7.584207 8.27368 8.9631 53 9.652627 10.3421 V(cm3/mol v (PR) 51 1.0997 382.4097 280.3229 177.3954 101.649 89.65217 84.43312 81.21061 78.91633 77.12994 75.69383 498.076 371.1377 ‘ 271.7131 175.101 109.1489! 95.9797 89.38151 85.037 81.82833 79.3 77.225 187 188 Table: Append"ix_3_-1 SLupereriticT-I Ethylene 80 F P(Psia) 500 600 700 800 900 1 000 1 100 1 200 1 300 1400 1 500 V(ft3llb) 0.31926 0.24619 0.19172 0.14789 0.1 1013 0.08024 0.06464 0.05726 0.05314 0.05049 0.0486 100 F P(Psia) 500 600 700 800 900 1 000 1 100 1 200 1 300 1400 1 500 V(ft3/lb) 0.34426 0.27016 0.21611 0.17436 0.14072 0.1 1315 0.09162 0.079677 0.06737 0.06131 0.0572 P(MPa) 3.447367 4. 13684 4. 82631 3 5.51 5787 6.20526 6.894733 7.584207 8.27368 8.9631 53 9.652627 1 0.3421 P(MPa) 3.447367 4. 1 3684 4.82631 3 5.51 5787 6.20526 6.894733 7.584207 8.27368 8.9631 53 9.652627 1 0.3421 V(cm3lmol V (PR) 559.1 395 431 . 1676 335. 7709 259.0088 1 92.8774 140.5292 1 13.208 1 00.2829 93. 06733 88.42622 85. 1 161 5 V(cm3/mol V (PR) 602.9236 473.1477 376.4666 305.3673 246.4515 196.1665 160.4597 139.5432 117.9892 107.3759 100.1779 545.9039T 419.54 326.25 252.5661 191.167 143.56 117.688 104.796 97.089I 91.807 87.86828. 589.77 461.53 368.65 297.77 241.692 196.747 161.9475 137.3667 121.2451 1 10.5723 103.106 189 TT‘abIe: Appendix 32 Supercritical Methane 0 F P(Psia) 400 500 600 700 800 900 1 000 1 100 1200 1400 1 600 1 800 120 F P(Psia) 400 500 600 700 800 900 1 000 1 100 1200 1400 1600 1800 V(fl3/|b) 0.70345 0.54934 0.44659 0.37323 0.31829 0.27569 0.2418 0.21431 0.1917 0.15722 0.13293 0.11561 V(ft3llb) 0.9341 0.74051 0.61 165 0.51978 0.45106 0.39777 0.3553 0.3207 0.29202 0.24737 0.21439 0.18923 P(M Pa) 2.757893 3.447367 4.13684 4.82631 3 5.515787 6.20526 6.894733 7.584207 8.27368 9.652627 1 1.03157 12.41052 P(MPa) 2.757893 3.447367 4.13684 4.826313 5.51 5787 6.20526 6.894733 7.584207 8.27368 9.652627 1 1.03157 12.41052 702.642 548.709 446.077 372.8013 317.9244 275.3733 241 .523 214.0638 191 .4798 157.0394 132.7773 1 15.4772 933.0271 739.6594 610.9475 519.183 450.5419 397.3131 354.8919 320.3316 291.6846 247.0859 . 214.1438 189.0126 V(cm3lmol v (PR) 602.89 542.64 503.81 476.417 455.814 439.614 426.457 415.5051 406.21 1 391.2046 379.533 370.137 V(cm3/mol V (PR) 868.8 738.296 657.42 602.41 8 562.4746 532.0424 508 488.47 472.25 446.74 427.5 41 2.38 1 190 mi. .23 83.2- 8686 235° 835... 5.35 0: «38 3% 8m 53. 83. ER? $58... 82.3 3.355 934.8 2.5 $.68 83 new 22 9:” 59.3- 983 $236 «323 «82.3. 8.8 ~33 «3 8m Sod so.” 883- $385 8°86 3836 «2.3.9. 8.3 9.8..” 3:. 8.3. new 8E 3.8%.? «$83 333° $28.0 88.8 $6.. 8.8... «2.» EN 53 B? 583 6233 388.0 898.6 383.3. e.g. 8&8 33 8a 83 23 «833 3386 858.0 9.886 9.5.33. 8.; .38 a: 8w «is. was. 6:6 9. ea 6.8 5.2820 5230 66.2 x .62“. >03“. 33. 633 P £22. £966 use: mOe<> Ema L. 3 BERN 2820 32 was. $3 "on. 293 v. Neon no» 80 a-” 52.2.2 3:3. 191 «Nomad 505 Ed rumored momumdm aovuwmd 333.0 3:. 3.0 mam—him wwnommd mvmm _.o.o «cam 36 Em 30. Pm womommd mmm..~o.o Nwmomod vmmoflv nonnmmd omvmmod «mammod ammodv 33.006 mmmowod mmmmwod 32.95 932.6 888.0 «8.4.8.0 $2.13 98 0.8 t 2.28 Emcee 9.6.8 3R6 ~83 nmdm omdm 05.3 mode ms. _.v mean 2.5 4023.20 4023.20 050... ”08.. $06 mad: vomé Nmémw mmmé 8.8.. mom...” «Yew word mmdwm mmod mmfimo rum... mOm<> was. wmwé x mdmw ._.03“. .9 £23 £28 050... «92> Ewe h 58.3. 3mm? .20sz $26 was. 9.? "on. 3mm... x 38 1.9. 2.2.2.. m-» 5.52:? 333. 193 00.0 500.0 500.0 000.0 505... 1050.? 00000 was. 40:“. ..00.0 500.0 000.0 000.0 505. .. 050. F 5000.0 was. >03". 0005.0? 53.0..- 5005?.5- 00000.0- 00055.0 030.50 0050? ...0 .25 .x. 0000000 000.000 0 90050.0 00030.0 00 .000 33000 0000050 h... 0000.00 5000 3.0 0000 3.0 5505 30 05.0 .00 550.000 00.0000 axm 36:2. 00.. .. .00 3.030 00:00 0050.00 00.00 0000000 0V0v000 0.8 £23 0000 _. .05 50000.05 00500.00 0003.50 50VVO. .0 303.5? ..000.vv 0' 5500.0 2.0.0 00.00 3.00 0.00 00.00 00.00 006v V0. 3. 402.020 40.20.20 050... 0100.. 000 .v v0.03. ..00.v 0100.. 000.0 50.000 F0 .0 00.000 0V0 .0 P0 .0 F0 000. .. 50.050 0.0. _. mr...)— mOn.<> .500 0000 .. .e 0&2 V006. v. 0.00.. 00.. 00 _. 00 _. 00 .. 05 .. 00 _. 00 _. v. ... "(Gm—20 non. no... 2.2.8:. 0.0 xwucona.‘ no.0: 194 Fable: Appendix 3-7 Ethylene TC= 282.4 K PC= MPa OMEGA= T density density Tr K calc exp 269 0.012189 0.012938 0.95255 255 0.014574 0.014641 0.902975 241 0.016372 0.015908 0.853399 225 0.018031 0.017078 0.796742 211 0.019248 0.017965 0.747167 °/o error -5.78915 -0.45762 2.916771 5.580279 7.141664 _A_ - Am.- % Error in Density 195 15 10 o- 5 4r- 0 -- -5 -— — Ethylene ‘::\ ‘\ -10 -- --— Methane ‘~\ \ ------ Propane \ - . _ . Argon A“ \ - - - - C02 ‘ 45 ’r % i i [L 0.7 0.75 0.8 0.85 0.9 0.95 T Figure Appendix 3-1: Error in the Peng- Robinson Calculation of Saturation Density LIST OF REFERENCES LIST OF REFERENCES Abraham, F. F .; Singh, Y. J. Chem. Phys. 1978, 68, 4767. Angus, 8.; Armstrong, B.; de Reuck, K. M.; Featherstone, W.; Gibson, M. R. Eds., ‘ International Union of Pure and Applied Chemistry; Butterworths: London, 1974, 39. Asselineau, L.; Bogdanic G.; Vidal, J ., Fluid Phase Equil. 1979, 3, 273. Barrer, R. M.; Robins, A. 3., Trans. Faraday Soc. 1951, 47, 773. Barton, S. S.; Dacey, J. R.; Quinn, D. F. In Fundamentals of Adsorption, Myers, A. L.; Belfort, G. Eds., Engineering Foundation: New York, 1983, 65. Berenyi, L. Z. Phys. Chem, 1923, 105, 55. Bienfait, M. In Phase Transitions in Surface Films, Dash, J. G.; Ruvalds, J. Eds., Plenum Press: New York, 1980, 29. Brennecke, J. F.; Eckert, C. A. In Supercritical Fluid Science and Technology; Johnston, K. P.; Penninger, M. L. Eds., ACS Symposium Series 406, American Chemical Society: Washington DC, 1989,14. Brennecke, J. F. Ph. D. Thesis, University of Notre Dame, 1989. Brennecke, J. F.; Tomasko, D. L.: Peshkin, J .; Eckert, C. A. Ind. Eng. Chem. Res. 1990, 29, 1682. Brunauer, S. The Adsorption of Gases and Vapors Vol. 1, Princeton University. Press, Princeton: New Jersey, 1945. Davis, H. T.; Scriven, L. E. In Advances in Chemical Physics, Prigogine, 1., Rice, S. A., Wiley: New York, 1982, 49, 357. 196 197 Debenedetti, P. G. Chem. Eng. Sci. 1987, 42, 2203. de Boer, J. H. The Dynamical Character of Adsorption, Clarendon Press: Oxford, 1953. Defay, L.; Prigogine, 1., Surface Tension and Adsorption, John Wiley, New York, 1966. Denbigh, K. The Principles of Chemical Equilibrium, 4th ed., Cambridge University Press: Cambridge, 1981. Dubinin, M. M. Proc. Acad. Sci. USSR. 1947, 55, 331. Dubinin, M. M. Progr. Surf Membrane Sci. 1975, 9, 1. Eckert, C. A., Ziger, D. H.; Johnston, K. P.; Ellison, T. K. J. Phys. Chem. 1986, 90, 2798. Evans, R. Advances in Physics 1979, 28, 143. Findenegg, G. H. In Fundamentals of Adsorption, Myers, A. L., Belfort, G. Eds., Engineering Foundation: New York, 1983, 207. Fisher, B. B., McMillan, W. G. J. Am. Chem. Soc., 1957, 79, 2969. ' Fisher, J .; Methfessel, M. Phys. Rev. A. 1980, 22, 2836. Gregg, S. J .; Sing, K. S. W. Adsorption, Surface Area and Porosity, Academic Press: London, 1982. Guggenheim, Thermodynamics, 5th ed., North-Holland Publishing: Amsterdam, 1967. Henderson, D. Fundamentals of Inhomogeneous Fluids; Marcel Dekker: New York, 1992. Hill, T. L. J. Chem. Phys. 1946, I4, 441. Hill, T. L. J. Chem. Phys. 1947, 15, 767. Hill, T. L. J. Chem. Phys. 1948, 16, 181. Hill, T. L. J. Phys. Chem. 1950, 54, 1186. Hill, T. L. J. Chem. Phys. 1951, I9, 261. Hill, T. L. In Advances in Catalysis, Frankenburg, W. G.; Rideal, E. K.; Komarewsky, V. I. Eds., Academic Press: New York, 1952, 4, 211. 198 Johnston, K. P.: Kim, S.; Combs, J. In Supercritical Fluid Science and Technology; Johnston, K. P.; Penninger, M. L. Eds. ACS Symp. Ser. 406, American Chemical Society: Washington DC, 1989. Kierlik E.; Rosinberg, M. L. Phys. Rev. A. 1990, 42, 3382. Kierlik E.; Rosinberg, M. L. Phys. Rev. A. 1991, 44, 5025. Kim, S.; Johnston, K. P. Ind. Eng. Chem. Res. 1987a, 26, 1206. Kim, S.; Johnston, K. P. AIChE J. 1987b, 33, 1603. Kung, W. C.; Scriven, L. E., Davis, H. T. Chem. Phys. 1990, 149, 141. Lee, L. L. Molecular Thermodynamics of Non-Ideal Fluids; Butterworths: Stoneham, MA, 1988. Lee, L. L.; Debenedetti, P. G.; Cochran, H. D. In Supercritical Fluid Technology: Reviews in Modern Theory and Applications, Bruno, T. J .; Ely, J. F. Eds., CRC Press: Boca Raton, F1, 1991, 193. Lennard-Jones, J. B.; Devonshire A. F. Proc. Royal Soc. A 1937, I63, 132. Maslan. F. D.; Altman, M.; Abereth, E. R J. Phys. Chem. 1953, 57, 106. McQuarrie, D. A. Statistical Mechanics; Harper & Row: New York, 1976. Myers, A. L.; Prausnitz, J. M. AIChE J. 1965, 11, 121. Nicholson, D.; Parsonage, N. G. Computer Simulation and Statistical Mechanics of Adsorption, Academic Press: New York, 1982. Peng, D. Y.; Robinson, D. B. Ind. Eng. Chem. Fund. 1976, 15, 59. Polyani, M. Trans. Faraday Soc. 1932, 28, 316. Powles, J. G.; Rickvayzen, G.; Williams M. L., Mol. Phys. 1988, 64, 33. Prausnitz, J. M. Proceedings of the 2nd International Conference, EFCE, DECHEMA: Berlin, 1980. ‘ Prausnitz, J. M., Liechtenthaler, R. N., de szedo, E. G. Molecular Thermodynamics of Fluid Phase Equilibria, Prentice-Hall, Englewood-Clifi‘s: New Jersey, 1986. Pyada, H. Masters Thesis, Michigan State University, 1994. 199 Radke C. J., Prausnitz, J. M. AIChE J. 1972, 18, 761. Rangarajan, B.; Lira, C. T. In Better Ceramics Through Chemistry V, Hampden-Smith, M. J ., Kemperer, W. G., Brinker, C. J. Eds. , Materials Research Society: Pittsburgh, PA, 1992, 559. Rangarajan, B.; Lira, C. T.; Subramanian, R. AIChE J. 1995, 41, 838. Reich, R., Ziegler, W. T., Rogers, K. A. Ind. Eng. Chem. Proc. Des. Dev.l980, 19, 336. Reid, R. C.; Prausnitz, J. M. Poling, B. The Properties of Gases and Liquids, 4th ed, McGraw-Hill: New York, 1987. Ross, S., Clark H., J. Am. Chem. Soc. 1954, 76, 4291. Ross, S.; Olivier, J. P. On Physical Adsorption, Interscience Publishers: New York, 1964. Rowlinson, J. S.; Widom, B. Molecular Theory of Capillarity, Clarendon Press: Oxford, 1982. Saam, W. F.; Ebner, C. Phys. Rev. A. 1978, 17, 1768. Sandler, S. L. Chemical and Engineering Thermoaynamics, 2nd ed., Wiley: New York, 1989. Snook, I. K.; Henderson, D. J. Chem. Phys. 1978, 68, 2134. Specovius J .; Findenegg, G. H. Ber. Bunsenges Phys. Chem. 1978, 82, 174. Specovius J .; Findenegg, G. H. Ber. Bunsenges Phys. Chem. 1980, 84, 690. Steele, W. A. In The Solid-Gas Interface Vol 1, Flood, E. A. Ed., Marcel Dekker: New York, 1967, 307. Steele, W. A. The Interaction of Gases with Solid Surfaces, Pergamon Press: Oxford, 1974. Strubinger, J. R.; Parcher, J. F. Anal. Chem. 1989, 61, 951. Subramanian, R.; Pyada H.; Lira, C. T. Ind. Eng. Chem. Res. (in press). Sullivan, D. E. Phys Rev. B. 1979, 20, 3991. Tan, C. -S.; Liou, D. -C. Ind. Eng. Chem. Res. 1988, 27, 988. 200 Tan, C. -S.; Liou, D. -C. Ind. Eng. Chem. Res. 1990, 29, 1412. Tan, 2.; Gubbins, K. E. J. Phys. Chem. 1990, 94, 6061. Tarazona, P.; Evans, R. M01. Phys. 1983, 48, 799. Teletzke, G. F.; Scriven, L. E.; Davis, H. T. J. Colloid Interface Sci. 1982a, 87, 550. Teletzke, G. F.; Scriven, L. E.; Davis, H. T. J. Chem. Phys. 1982b, 77, 5794. Telo de Gama, M. M.; Evans, R. Mol. Phys. 1983, 48, 687. Tjatopolous, G. J .; Feke, D. L.; Adin Mann, Jr., J. J. Phys. Chem. 1988, 192, 4006. Valenzuela, D. P., Myers, A. L. Adsorption Data Handbook, Prentice-Hall, Englewood- Cliffs: New Jersey, 1989. Vanderlick, T. K.; Scriven, L. B.; Davis, H. T. J. Chem. Phys. 1989, 90, 2422. van Megen, B.; Snook, I. K. Mol. Phys. 1982, 45, 629. van Megen, B.; Snook, I. K. Mol. Phys. 1985, 54, 741. Vera, J. H.; Prausnitz, J. M. Chem. Eng. J. 1972, 3, 1. Wu, R. -S.; Lee, L. L.; Cochran, H. D. Ind. Eng. Chem. Res. 1990, 29, 977. Yang, R. T. Gas Separation by Adsorption Processes, Butterworths, Stoneham, MA, 1987. Young, D. M.; Crowell, A. D. Physical Adsorption of Gases, Butterworths, London, 1962. "llllllllllllllll