A GENERALIZED THEORY OF SORPTTON PHENOMENA IN BIOLOGICAL MATERIALS Thesis for the Degree of Ph. D. MMBHIGAN STATE UNIVERSITY PATRTCK OBI NGODDY 196‘9 THEQS This is to certify that the thesis entitled A GENERALIZED THEORY OF SORPTION PHENOMENA IN BIOLOGICAL MATERIALS presented by Patrick Obi Ngoddy has been accepted towards fulfillment of the requirements for Ph.D. degree in Agr. Eng. 4 Date sent. 241 1969 0-169 /./." ’ //. / /.7 A 1:, g ml. LIB P 4R Y Mic‘mg. :1 State University I Lla-.bhuniu=-E‘i-'=i-III=IE; It ‘ ’0’)». Ketlltmtlll ABSTRACT A GENERALIZED THEORY OF SORPTION PHENOMENA IN BIOLOGICAL MATERIALS by Patrick Obi Ngoddy A moisture sorption model unifying three basic concepts in adsorption, is developed and verified for biological materials. The B-E-T and Capillary Condensation theories are combined into an integral isotherm equation for porous biological materials. In order to solve the integral equation explicitly, it was necessary to (a) characterize the physical structure of sorbing bio-materials mathematically, and (b) to derive an explicit density function for water sorbed by biological materials, utilizing the general frame- work of the Adsorption Potential theory in conjunction with the fundamental laws of thermodynamics. Application of the moisture adsorption model to adsorption data of a wide variety of biological materials led to the following Specific conclusions: 1. The pores of biological materials are best modeled as inter- connected spheroidal ink-bottles. 2. The numerical distribution of pore-size in biological materials can be described by a power law type analytical function. 3. The density of water adsorbed by biological materials may be expressed by the semi-theoretical relation: = * ° p/po u (AHst/AHSt) where p is the density associated with the isosteric heat of 8' o .. ‘ -’mw'+—fl 1..“ m ads< isos defj 1501 A t' of SW SUC 5. Patrick Obi Ngoddy adsorption AHst’ po and AH;t are the corresponding density and isosteric heat values at saturation, and u* is empirically defined for the class of bio-materials of interest. The three basic concepts in adsorption can be unified into an isotherm equation of the type: Ma = 355 (Zn-Kn) l: and n are characteristic parameters of the physical structure of the sorbent, and Z and A are functions of relative vapor pressure. A theory of sorption hysteresis based on the superposition of the so-called "capillary condensation" hysteresis and "swelling fatigue" hysteresis is deve10ped and used to predict successfully the desorption isotherms of whole corn kernels. Approvediqal Zfl/gZ/v/Zt COIL/1114 fiajor ProfessorJ(//.7 X4 ((7 Approved CUM M M Department Chairman A GENERALIZED THEORY OF SORPTION PHENOMENA IN BIOLOGICAL MATERIALS By Patrick Obi Ngoddy A THESIS Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Agricultural Engineering 1969 Ge/ 75? 4024-170 To The pe0ple of BIAFRA at their moment of supreme anguish. ACKNOWLEDGEMENTS It gives the author special pleasure to acknowledge with sincere gratitude, the indispensable help and guidance of Dr. Fredrick W. Bakker-Arkema (Agricultural Engineering), a truly inspiring teacher and counselor, whose attitude and friendship made this investigation a very rewarding and delightful experience. The personal interest and unfailing assistance of Dr. J. B. Kinsinger (Chairman, Chemistry) who guided the author's study of the fascinating field of Surface Chemistry is gratefully acknowledged. Sincere appreciation is extended to Dr. W. C. Bickert (Agri- cultural Engineering) and Dr. D. R. Heldman (Agricultural Engineering and Food Science) for serving on the author's guidance committee. The author is indebted to Dr. Carl W. Hall (Chairman, Agri- cultural Engineering) and his staff for making available the graduate assistantship and for their kind c00peration and help in innumerable other ways. Grateful acknowledgement is due Dr. Bakker's group of graduate students, David Farmer, Gonzalo Roa, John Rosenau, and David Thompson, whose lively participation in several brain storming sessions helped to crystallize many of the concepts presented in this dissertation. It seems apprOpriate to recognize here, members of the author's family: his parents, Mr. and Mrs. H. 0. Ngoddy and sister, Mrs. R. A. Akudu, who were directly reSponsible for his early education; his wife, Omogo, without whose understanding and kind support this study could not have been accomplished. Finally, thanks go to the members of the M.S.U. Glass ShOp for their unfailing assistance, to Mr. M. P. Palnitkar for his assistance in the laboratory experiments, and to Mrs. Nancy Hodge for her extra- ordinary patience and diligence in typing the manuscript. 7':~k*~k‘k*k*7‘r~k 2.4 2.5 TABLE OF CONTENTS Page INTRODUCTION . . . . . . . . . . . . . . 1 1.1 General Remarks . . . . . . . . . . . . 1 1.2 Hygrosc0pic Phenomena - Moisture Retention and Transport in Bio-Materials . . . . . . . 2 1.3 The Nature of Adsorption . . . . . . . . 3 1.4 Adsorption as a Function of the Nature of the Adsorbent O O O O O U C O O O O O O O 6 1.5 Adsorption as a Function of the Nature of the Adsorbate - The Water Molecule . . . . . . . 8 1.6 Review of Related Literature . . . . . . . 12 1.7 Statement of the Problem . . . . . . . . . 28 1.8 Objectives . . . . . . . . . . . . . 29 THEORETICAL CONSIDERATIONS . . . . . . . . . . 30 2.1 Introduction . . . . . . . . . . . . . 30 2.2 A Physical Model of Water Adsorption by Bio- Materials . . . . . . . . . . . . . 30 2.3 Generalization of Molecular and Capillary Adsorbed Moisture . . . . . . . . . . . 33 (a) Pore Characterization - defining the size of a pore filled by Molecular and Capillary Adsorption . . . . . . . . . . . . 33 (b) The thickness of the adsorbed multi-layer . 37 (c) The Isotherm Equation . . . . . . . . 41 2.4 Adsorption Compression on Adsorbing Bio—Materials 48 2.5 Sorption Hysteresis in Biological Materials . . 68 (a) Capillary Condensation Hysteresis . . . . 69 (b) Swelling Fatigue Hysteresis . . . . . . 72 (c) Double Superposition of "Capillary Condensa- tion" and "Swelling Fatigue" - Hysteresis . 76 i III. IV. EXPERIMENTAL . . . . . . . . . . . . . 3.1 Product Preparation . . . . . . . . . 3.2 Determination of Equilibrium Moisture Content (6) (b) The Adsorption System . . . . . . Procedure . . . . . . . . . . RESULTS AND DISCUSSION . . . . . . . . . 4.1 The Determination of Pore Structure from Water Sorption Isotherms-Verification of a Pore- Size Distribution Function of the Power Law Type for Bio—Materials . . . . . . 4.2 Experimental Results . . . . . . . . 4.3 Verification of the Derived Density Function for Sorbed Water . . . . . . . . . 4.4 Verification of the Derived Isotherm Equation . (a) (b) (C) (d) 4.5 Verification of the Preposed "Capillary Condensa- tion Theoretical Isotherms . . . . . . . n as the Primary Characteristic Parameter of the Pore Structure . . . . . . . g as the Secondary Characteristic Parameter of the Pore Structure . . . . . . Prediction of the Adsorption Isotherms of Certain Biological Materials with the Generalized Isotherm Equation . . . - Swelling Fatigue" Superposition Theory of Sorption Hysteresis in Biological Materials (8) SUMMARY AN SUGGESTION REFERENCES APPENDIX The Parameter, w, as a factor of pore geometry . . . . . . . . . . D CONCLUSIONS . . . . . . . . . S FOR FURTHER STUDY . . . . . . . ii Page 77 77 78 78 83 85 85 103 111 119 119 121 122 123 144 147 149 152 154 163 Table 2.1 2,2 2.3 2.4 4.1.1 A.l.2(a) 4.1.2(b) 4.1.3 4.2.1 4.2.2 4-3.1 4.3.2 Table 2.1 2.2 2.3 2.4 4.1.1 4.1.2(a) 4.1.2(b) 4.1.3 4.2.1 4.2.2 4.3.1 4.3.2 LIST OF TABLES The function u* as determined from the empirical data of Stamm (1938) for spruce wood at 25°C. . . . . . . . . The reduced function p/po = u*(AH/AH°) for corn. Based on the desorption isotherm (4°C) data of Rodriguez-Arias (1956) . . . . . . . The reduced function p/p = u*(AH/AH°) for cotton. Based on the Adsorption isotherm (10°C) data of Urquhart and Williams (1924) . . The reduced function p/p = u*(AH/AH°) (for pre- cooked freeze-dried beef powder based on adsorp- tion isotherm (10°C) data obtained in the experi- mental part of this study . . . . . . . . Illustrative input data for pre-cooked beef powder freeze-dried at 105°F platen temperature Pore-Size Distribution for pre-cooked beef powder freeze-dried at 105°F platen temperature. Calcu— lated by a modified Cranston and Inkley scheme, and based on 10°C isotherm. A cylindrical pore model is assumed . . . . . . . . . . Pore-Size Distribution for pre-cooked beef powder freeze-dried at 105°F platen temperature. Calcu- lated by a modified Cranston and Inkley scheme, and based on 10°C isotherm. A spheroidal Ink Bottle pore model is assumed . . . . . . . Summary of the characteristic parameters of pore structure as determined from pore-size distri- bution plots . . . . . . . . . . . . Equilibrium moisture content of pre-cooked freeze-dried ground beef. Adsorption . . . . Equilibrium moisture content of pre-cooked freeze-dried ground beef. Desorption . . . . Verification table for the density of sorbed water. Product = Pre-cooked freeze-dried beef powder . . . . . . . . . . . . . Verification table for the density of sorbed water. Product = Soda boiled cotton . . . . . iii Page 60 61 63 64 89 91 94 101 104 106 114 115 Tab‘ L‘ ‘\ 0 £\ L\ Table 4.3.3 4.3.4 4.4.1 4.2.2 4.4.3 4.4.4 4.4.5 4.5.1 4.5.2 Verification table for the density of sorbed water. Product = spruce wood . . . . Verification table for the density of sorbed water. Product = Corn kernels . Calculation of Theoretical Isotherm (a) Pre-cooked freeze-dried beef powder (10°C) (b) Pre-cooked freeze-dried beef powder (37.7°C) Calculation of Theoretical Isotherm (a) Raw freeze-dried beef slices (10°C) (b) Raw freeze—dried beef slices (40°C) Calculation of Theoretical Isotherm (a) Whole Corn Kernels (22°C). . . . (b) Whole Corn Kernels (50°C). . . . Calculation of Theoretical Isotherm (a) Wood (20°C) . . . . . . . . (b) Wood (60°C) . . . . . . . . Calculation of Theoretical Isotherm (a) Cotton (10°C). . . . . . (b) Cotton (30°C). . . . . Work sheet for the calculation of desorption isotherm of whole corn kernels at 22°C . Work sheet for the calculation of desorption isotherm of whole corn kernels at 50°C . iv Page 116 117 125 126 128 129 131 132 134 135 137 138 145 146 Figure 1.1 2.1 2.2 2.3 204 2.5 2.6 2.7 2.8 2.9 2.10 2.11 2.12 3.1 3.2 LIST OF FIGURES Molecule of water vapor . . . . . . . . Pictorial representation of an organic tissue as a random network of small, irregular pores. The diameter of one pore is magnified for illustrative purposes . . . . . . . . Cylindrical capillary of radius, R, showing the "inside annular tube", r, which may be given either by the Cohan or Kelvin equation . . . Capillary menisci (a) Hemispherical (Kelvin) meniscus . . . (b) Cylindrical (Cohan) meniscus . . . . Packing of adsorbed molecules in successive molecular layers, to illustrate variable stacking configurations . . . . . . . . Plot of X/X against P/P . . . . . . . m o Shapes of capillaries (a) Cylindrical capillary . . . . . . . (b) Interconnected spheroidal "Ink Bottles" . Polanyi's iSOpotential contours . . . . Diagram showing two thermodynamically identical systems (a) Contains an active sorbent . . . . . (b) Contains the pure sorbent . . . . . (0/00) and (AH/AH°) plotted against X/Xm . . Plot of the function u* as obtained from the empirical data of Stamm (1938) . . . . . Generalized plot of the function, u*, as obtained from the empirical data of Stamm (1938) . . Plot of p/po against X/XL . . . . . Adsorption apparatus . . . . . . . . Diagram showing the principle of the Cahn- Electrobalance . . . . . . . . . . Page 10 31 36 36 36 38 40 44 44 49 53 53 65 65 66 67 79 80 Figure 3.3 4.1 4.3 4.4 4.6 4.7 4.8 4.9 4.10 4.11 4.12 4.13 4.14 4.15 (a) Close-up view of the electrobalance in the vacuum line . . . . . . . . . . (b) Close-up view of the Control unit and recorder outside the experimental chamber . . Pore-Size distribution plot for corn at 4°C, 15.5°C, 30°C isotherm temperatures . . . . . . Pore-Size distribution plot for cotton at 10°C, 20°C and 30°C isotherm temperatures . . . . . Pore-Size distribution plot for wood at 20°C, 60°C and 100°C isotherm temperatures . . . . . Pore-Size distribution plots for pre-cooked freeze-dried beef powder at 10°C, 22.2°C and 37.7°C isotherm temperatures . . . . . . . Adsorption isotherms for pre-cooked freeze-dried beef pOWder (10°C ' 37.7°C) o o O o o o o o Desorption isotherms for pre-cooked freeze-dried beef powder (10°C - 37.7°C) . . . . . . . (a) Plot showing sorption hysteresis in pre- cooked freeze-dried beef powder (10°C and 37.7°C) (b) Plot showing sorption hysteresis in pre-cooked freeze-dried beef powder (22.2°C) . . . . . Vertical section of stretched out specific surface, showing the multi-layer matrix . . . . (a) Theoretical isotherms, Pm 0.008 PO . . . (b) Theoretical isotherms, Pm = 0.1 PC . . . . Comparison of experimental adsorption isotherms with calculated isotherms for pre-cooked freeze- dried beef powder . . . . . . . . . . Comparison of experimental adsorption isotherms with calculated isotherms for raw-freeze-dried beef slices . . . . . . . . . . . . Comparison of experimental adsorption isotherms with calculated isotherms for corn . . . . . . Comparison of experimental adsorption isotherms with calculated isotherms for wood . . . . . . Comparison of experimental adsorption isotherms with calculated isotherms for cotton . . . . . Comparison of experimental desorption isotherms with calculated isotherms for corn . . . . . vi Page 81 81 97 98 99 100 105 107 108 109 112 120 120 127 130 133 136 139 148 d.b. AH .Q st st AH, A'H AH ,A'H m m AF 12 L(R) NOTATION average diameter of sorbate molecule, [A°] factor of pore geometry dry basis [calories] enthalpy, isosteric heat of sorption, [calories/mole} enthalpy change, [calories/gram] molar enthalpy change elastic modulus, [psi] surface free energy bulk modulus, [psi] pore-size distribution parameter, eqn. (2.3.16) parameter defined by eqn. (4.1.5) length of pore of radius, R, [A°] molecular weights of sorbate, [grams] specific adsorbed mass [grams] specific desorbed mass equilibrium moisture content, [M.C.d.b.] moisture content number of sorbate molecules, number of adsorbed multi—layers vii AS,A'S AS ,A'S m m vapor pressure vapor pressure corresponding to the mono-layer capacity adsorption vapor pressure desorption vapor pressure swelling pressure, eqn. (2.5.5) pore radius, [A°] universal gas constant radius of curvature of meniscus during adsorption radius of curvature of meniscus during desorption [A°] generalized radius of meniscus [A°] Cohan radius [A°] Kelvin radius [A°] parameter defined by eqn. (4.1.4) entrOpy, [Cal./degreej o 2 B-E-T surface area (A ) molar entrOpy, [Ca1./degree-mole] entrOpy change molar entrOpy change N(R) 12 Av number of pores of size, R used as a subscript or superscript implies saturation conditions specific adsorbed volume, eqn. (2.3.13) [cc] volume, [cc] molar volume [cc] volume of pore of radius, R finite volumetric step volumetric hydro- expansion in chapter I used as relative vapor pressure; elsewhere used as specific adsorbed mass Specific adsorbed mass at the mono-layer capacity specific adsorbed mass at hygrosc0pic or fiber saturation point variable, defined by eqn. (2.3.23) identification for a com- ponent of a thermodynamic system coefficient of linear hydro—expansion, % M.C.d.b.] surface tension of con- densed sorbate normal stress, [psi] viii [in./in. 12 U) if“ m (R) w2(R) surface area pores associated with a given adsorption step. [(A°) 3 temperature, [°K] thickness of adsorbed multi-layer, [A° J variable defined by eqn. (2.3.25) variable defined by eqn. (2.3.26) factor of pore geometry empirical constant of the Halsey eqn. (2.3.5) average thickness of a single layer of adsorbed molecules, [A°] pore-size distribu- tion parameter, eqn. (2.3.16) constant defined by eqn. (2.3.24) primary characteristic parameter of pore struc- ture, eqn. (2.3.20) secondary character- istic parameter of pore structure, eqn. (2.3.19) function of pore geometry pore-size distribution function Poisson's ratio luv-936.331.» I. IRWIN“; 1» .. area which one adsorbed V,¢ molecule will occupy (A°)2 n density of sorbed water, [Ema/0C0] 8 function defined by eqn. (2.4.26) chemical potential K ix contact angle two dimensional spreading pressure in chapter I fraction of surface covered by mono- molecular adsorption factor of pore geometry I. INTRODUCTION 1.1 General Remarks The growing socio-economic importance of food and fibres together with the complexity of the technology for their production, handling, processing, preservation and distribution has brought about a new and exacting emphasis among Agricultural and Food Engineers on the study of BIOLOGICAL MATERIALS SCIENCE. The primary intent of this endeavor is to provide higher quality food and fibre pro- ducts more economically, while the methodology consists of the basic engineering approach to problems involving the handling and proces- sing of biological materials. It appears logical that the physical laws governing the response of biological materials to handling and processing must be well understood in order to Optimize harvesting, processing, handling and storage systems. It is now well established that the moisture content of bio- logical products exerts a profound influence on their mechanical, physical, chemical and enzymic behaviors. This recognition has, in the main, provided the stimulus for the large body of scientific work to be found on the subject. The work, with a few limited exceptions, has been primarily experimental in nature. In consequence, while there exists a vast quantity of confirmatory data on the relation- ship between biological products and those molecules of water asso- ciated with them, the exact nature of water in food, and even such fundamental datum as the adsorption isotherm itself has not been satisfactorily derived from a theoretical point of view. The cause of the sigmoid shape of this isotherm and of the distinctively pronounced hysteresis effect must still be considered the subject of conjecture rather than the consequence of an established theory. A large body of isolated data still exists which to a greater or lesser degree has refused to fit any general treatment. Yet, in view of its all pervading influence, it is fundamental to the entire field of biological materials science, that the laws which underlie the hygroscOpic phenomena in biological materials be well understood and particularly that these laws be rendered more con- cise, understandable and useable by expressing them in more pre- cise mathematical terms. The present work was undertaken in the attempt to weld some of the data - old and new - into a coordinated theory of adsorption for biological materials. 1.2 Hygroscopic Phenomena — Moisture Retention and TranSport in Bio-Materials Kuprianoff (1958) in his treatment of bound water in foods, suggests that moisture in biological materials may exist as free moisture, chemically bound moisture, and adsorbed moisture. In appraising the relative contribution of each type of moisture to the overall role of water in food, he concludes that adsorbed moisture is by far the most important category of water in biological products. The extensive body of scientific work on the subject amply vindicates the point that problems associated with moisture transport and reten- tion in bio-materials above the so-called hygrosc0pic point lend themselves to relatively easy theoretical treatment. The hygrosc0pic point has been defined by Lewis (1921) as the product moisture content correSponding to a relative humidity of 100 percent. Above this point, the physical laws of osmosis, liquid diffusion and capil- larity govern the motion or retention of water molecules in organic materials. These laws are coordinated with remarkable success (Lykov, 1955) into a moisture transport theory. This theory has been variously used (Van Arsdel and Capley, 1963) to reconstruct the so-called constant rate drying period which characterizes the drying of biological materials above the hygrosc0pic point. In contradistinction with the so-called "free or removable" moisture discussed above, water associated with biological products below the hygrosc0pic point is subject to intense surface phenomena called sorption. This explains why this type of moisture is called adsorbed moisture. The removal of this type of moisture is character- ized by a falling rate drying period. The motion and retention of adsorbed moisture in biological materials have eluded rigorous theoretical treatment for a long time. The reason why this is the case will become apparent in subsequent discussions in this chapter. 1.3 The Nature of Adsorption When a solid is exposed to a vapor at a definite pressure, the concentration of the vapor is higher at the solid - vapor interface than it is in the vaporous atmosphere. This is the phenomenon called adsorption. It is not confined to solid-gas interface alone, but must be expected to occur in gas-liquid, liquid-solid, liquid-liquid and even in some circumstances solid-solid interactions (Brunauer, 1945). Adsorption is described phenomenologically in terms of an empirical general adsorption function, m = f(P,T) where m is the amount of vapor adsorbed under conditions specified by pressure P and temperature T. As a matter of experimental convenience the relationship between the amount of vapor adsorbed by a solid and vapor pressure is often represented by the so-called moisture equilibrium isotherm, m = fT(P). This isothermal function can be generated for several temperatures making it thereby possible for alternative plots of the data as isobars, m = fp(T) or isosteres, P = fm(T) to be develOped. While a major body of experimental and theoretical work has focused on the precise reconstruction of the adsorption isotherm, the isostere has almost exclusively been used in conjunction with the Clausius-Clapeyron equation to obtain the heats of adsorption. It is customary to divide adsorption into two broad classes, namely physisorption and chemisorption. Physical adsorption as the former is alternately called, takes place relatively rapidly, and, apart from hysteresis, the process is reversible. It is further supposed that adsorption is induced by the same type of relatively nonspecific intermolecular forces responsible for condensation phenomena. Thus, in physical adsorption, the heat of adsorption should be in the range of the heats of condensation. Physical adsorption is important only for gases below their critical tempera- tures, i.e., for vapors (Adamson, 1967). Chemisorption, on the other hand, may occur below or above the critical temperature of the sor- bate. It may be slow or rapid. It is qualitatively distinguishable from physisorption in that the chemical specificity is higher and the energy of adsorption so large to suggest the presence of an activation energy necessary for full chemical bonding. While the probability of chemisorption playing at least some role has been considered by some investigators (Rodriquez-Arias, 1956), it is now generally agreed that is is adsorption of the van der Waal's type which is almost wholly in effect in the adsorption of water vapor by biological products. Even though no definitive concensus is discernible from the literature on the state of the solid-adsorbate complex, certain directive conclusions can be drawn regarding moisture retention in biological materials by a careful comparison of available experimental data and existing adsorption theory. While such conclusions are at best tentative and must be updated as the frontiers of uncertainty in surface phenomena shrink, they remain essential if our present modelistic conceptualization of adsorption in biological materials is to continue its steady advance. Bearing this in mind, observations regarding moisture retention by biological materials can be summarized as follows: their predominant sigmoid shape together with their pro- nounced hysteretic behavior distinguish the moisture equilibrium iso- therms of biological products as belonging to the type II (sometimes type IV) classification of the Brunauer type. If the combined con- sequences of (a) the B-E-T theory of multimolecular adsorption, (b) Polanyi's adsorption potential theory, and (c) Zsigmondy's capillary condensation theory are now admitted into the picture, it becomes rather obvious that moisture is held in organic materials by inter- molecular and capillary adsorption forces. ...-ss.|. hilt . 1.1.. fill) 441% Molecular adsorption occurs when the water molecules adhere to the pore walls as result of the divergent force field at the surface of the cell walls. When moisture is adsorbed by organic materials, it is compressed due to strong intermolecular forces of attraction. The adsorption compression is highest at low rela- tive humidities, so that there is a net decrease in the volume of the water-solid aggregate. As layers of water molecules on the cell wall increase with rising vapor pressure, the force of attrac- tion is decreased and the resultant adsorbate compression is corres- pondingly diminished. This simplified picture of molecular adsorp- tion is essentially in harmony with the "hydrogen bonding" theory of adsorption as discussed by Ward (1962). Capillary adsorption occurs when voids in the cellular struc— ture of a material are of a size large enough to hold water in liquid form, under reduced vapor pressure, by the forces of surface tension. The tensions originating in the capillary water are transmitted to the capillary walls and there produce stresses within the cell walls. These stresses combined with swelling stresses culminate in what Hammerle (1968) termed the HYDRO-STRESS in biological materials. It will be shown later in this work that the hydro-stress can be used as a basis for the quantitative estimation of hysteresis in biological materials. 1.4 Adsorption as a Function of the Nature of the Adsorbent The quantity of the vapor adsorbed by a given weight of adsor- bent varies greatly from one adsorbent to another. Since, as has been amply discussed in the preceding sections, adsorption is a surface phenomena, this variation must be due to the active Operation of at least two factors: the area of the interface and the specific adsorbing prOperties of the substance per unit area. Because the present interest lies in the class of organic plant or animal materials, it must be quickly recognized that these are character- ized by a great complexity and heterogeneity in physical structure. Each being an assemblage of strongly hydrated high-molecular-weight compounds, mostly belonging to the classes called proteins, carbo- hydrates, and lipids. Complete quantitative representation of even a single one of these systems is not possible. Consequently, the average or combined behavior of a large number of individual units or pieces must be dealt with in practice. Bearing this limitation in mind, both plant and animal tissues can be considered on the microsc0pic level as being funda- mentally cellular in their natural state. These fundamental building blocks are interposed with a complex labyrinth of passageways making them microsc0pically intricate porous bodies (Babbitt, 1942). While it is impossible to estimate from a priori considerations the poten- tial surface area that plays a part in adsorption phenomena, the present state of the art has made possible the rough estimation of the relative pore size distributions in such porous complexes. If this tool is used with caution, it is now entirely within the realms of possibility to characterize the physical structure of biological materials mathematically in terms of a pore-size distribution func- tion and an idealized pore geometry. It is the thesis of this study that if the structure of an adsorbing biological material can be so mathematically characterized, its equilibrium isotherm shape can be predicted accordingly. The details of this structural characterization will be pursued in a subsequent chapter. 1.5 Adsorption as a Function of the Nature of the Adsorbate - the Water Molecule As has become evident from the foregoing discussions, the adsorbability of water vapor by organic tissues must depend in a very high degree upon the nature of water itself. While within the context of a lumped or averaged treatment of the subject, water must necessarily be conceived as an ubiquitous and fundamental part of the structure of biological products (Kuprianoff, 1958; Barkas, 1953), cognizance must here be taken of the fact that scientists and engineers often find it convenient to assume a double super- position effect in defining the various properties of water-binding biological products (see for example Hammerle, 1968). It becomes therefore expedient to consider at this point, the nature and pro- perties of the water molecule. Water is, in many respects, a unique compound. Chemical reactions and physical interactions in which it participates on the molecular scale influence every gross characteristic of organic materials. Except for a slight natural ionization which leads to the formation of minute amounts of hydrogen and hydroxyl ions, pure water consists of molecules made up of two atoms of hydrogen and one atom of oxygen. These molecules may be aggregated by weak forces into quasi-crystalline combinations whose size and form depend upon the physical conditions prevailing at the time. Thermochemical data show that the heat of formation of gaseous water from hydrogen and oxygen gases is 57.8 K Ca1./Mole. Since the heats of dissociation of hydrogen and oxygen gases are 103.4 and 118.2 K Cal./Mole., reSpectively, the bond energy for each O-H bond is one-half of the sum: 57.8 + 103.4 + 59.1 or 110.2 K Ca1./Mole. The chemical bonds are formed by completion of two electron pairs, each of the two unpaired electrons of the oxygen atom associating with the electron of a hydrogen atom. Four of the remaining six electrons of oxygen are much farther apart from the nucleus than the other two. These four plus the four electrons involved in the chemical bonds tend to form four pairs that are as far apart as they can be while still attracted by the oxygen nucleus. They form the corners of an imaginary tetrahedron. The water molecule is schematically represented in Figure 1.1. The distance of the boundary shell from the center of the atom is equal to the so-called van der Waal's radius. The equivalent diameter of the molecule as determined from the density of ice is about 3.2 A. The cross-sectional area of the water molecule is given either as 14.8 (A)? corresponding to its chemisorption as O-H groups or 10.6 (A)2, corresponding to physical adsorption as H20 molecule along its longest axis (Gregg and Sing, 1967). These are important quan- tities which will be used extensively later in this work. Water vapor exhibits phenomena which can be most satisfactorily explained by assuming that the molecules involved therein contain 0' = 104° 31' Figure 1.1. Molecule of water vapor. 10 rigid, or nearly rigid, electrical dipoles. The distance between the dipoles being significantly less than the diameter of the molecule. The magnitude of these moments can be obtained from a study of the dielectric constants (Dorsey, 1940). The unique prOperties of water with respect to molecules of similar size, such as NH3, H23 and CH3OH, have been attributed (Bernal and Fowler, 1933) not only to its dipole character, but also to the geometrical structure of the molecule. In the present context, the later char- acteristic together with the polarizability of the water molecule (i.e., the distortion in the molecular charge distribution and the resultant distortion of molecular geometry due to external force fields) may in fact be more important than the dipole qualities (Adamson, 1967). In the vapor, the molecules are separated by relatively great distances and are moving at high velocities. The trans- lational energy of these molecules is so great that the van der Waal forces are inadequate to hold them together when they collide. Thus the vapor expands and exerts pressure in conformity with the kinetic theory of gases and in the first approximation behave in accordance with the ideal gas laws. In the liquid the molecules are held in intimate contact with one another by combined intermolecular forces. Each molecule in liquid water occupies a volume of 29.7(A)3 which indicates a porosity between molecules of about 36.7% (Matz, 1965). The nature of the bonding forces is not well understood. Some contribution is probably made by the weak forces of the van der Waal's type (with 11 bond energies of about 5 K Cal./Mole.) and must be, to a consider- able extent, nondirectional, since the molecules are free to move in the liquid. The magnitude of the attractive forces between the molecules must be relatively large, since the vapor pressure of the liquid is negligible compared to the pressure which is implied by the gas laws. In organic products hydrogen bonding exists between water molecules, between water and small and large molecules and ions, and as the direct link between these latter micromolecular components (Ward, 1962). There is, therefore, competition between the compon- ents for the available hydrogen bonding sites. The rate at which water may break or promote particular intermolecular hydrogen bonds, and the extent to which this can be carried out, together with the converse process of displacement of water from bonding sites, affect all practical dehydration and rehydration processes. Coulson (1959) reduced the intermolecular forces to (a) electrostatic interactions (b) diSpersion forces (c) repulsive forces and (d) delocalization interaction. Under the combined influence of these forces the water molecule in its adsorbed state appears to have undergone a definite thermodynamic transformation (Ewing and Spurway, 1930). While there exists no satisfactory theory of the liquid state, the detailed treatment of the adsorbate-adsorbent complex is even farther away. 1.6 Review of Related Literature In view of the extensiveness of the subject, it is not intended here to give complete coverage of the voluminous literature 12 on the general subject of gas adsorption. Instead, the principal models or theories which have been advanced to account for the sorptive behavior of biological materials or from which such theories are directly or indirectly derivable, will be taken up partly for their own sakes and partly as a means of introducing a new general- ized model to be presented in a later chapter. In order to expedite the task in this section, it is necessary to outline some criteria against which the works to be examined can be evaluated. As was previously mentioned, the adsorption isotherm is not only the most convenient form in which to obtain and plot experimental data, it is also the form in which theoretical treat- ments are most easily develOped. The first demand of a theory of adsorption then, is that it gives an experimentally correct adsorp- tion isotherm. Adamson (1967), however, points out that this is a necessary but insufficient test of the validity of the premises underlying a theory. Since quite differently based models have been found to yield equations which are experimentally indistinguishable and even algebraically identical, he suggests that data on the heats and possibly entrOpies of adsorption be used as a more discriminating test of an adsorption model. The characteristic hysteresis observed in adsorption by porous systems provides a third criterion of evaluation in the case of porous biological materials. This demands that the hysteresis phenomenon be both qualitatively and quantita- tively derivable from a given adsorption model. 13 Labuza (1968) in a recent review of the subject, points out that theoretical treatment of sorption has primarily come under three modelistic frameworks, namely: 1. The Kinetic concept of Langmuir; 2. Polanyi's Adsorption potential theory; and 3. Zsigmondy's capillary condensation theory. The Kinetic Concept: Langmuir (1918) preposed the classical kinetic model of adsorption based on his belief that adsorption was in essence, induced by unbalanced chemical forces on the surface of crystals leading to a unimolecular layer of the adsorbate. Assuming that (a) adsorbed molecules are localized, (b) colliding adsorbate molecules are reflected elastically, and (c) the heat of adsorption for every adsorbate molecule striking the bare surface of a solid adsorbent is the same and equal to the heat of vaporization, Langmuir equated the rate of evaporation to the rate of condensa- tion at the surface of the adsorbing solid under conditions of dynamic equilibrium. The resultant isotherm equation is of the form: P/v = 1/b vm + P/vm [1.1] where v is the volume of gas adsorbed isothermally at vapor pressure P; vm is the volume adsorbed at the monolayer point; and b is a con- stant of adsorption dependent on both the heat of adsorption as well as the isotherm temperature. In Spite of the limitations inherent in its simplifying assumptions, the Langmuir isotherm equation is perhaps the most important single equation in the field of adsorption, serving in many cases as the starting point in the derivation of other equations. 14 Type I adsorption isotherms are best interpreted in terms of Langmuir's equation and accordingly, are sometimes referred to as Langmuir isotherms. While Langmuir's pioneering adsorption model may be con- sidered a perfectly acceptable one at very low P/Po values, its greatest merit lies in the fact that it forms the basis of the more universally accepted B-E-T theory of multimolecular adsorption. This theory has, with varying degrees of success, been used to account for the occurence of types II, III, IV and even V adsorp- tion isotherms of the Brunauer classification. Brunauer, Emmett and Teller (1938) extended Langmuirs kinetical approach to multi- layer adsorption. The basic assumptions in the B-E-T model are: (a) the Langmuir equation applies to each layer, (b) for the first layer the heat of adsorption, q, may take a unique value, whereas for all succeeding layers, it is equal to qv, the heat of vaporiza- tion of the liquid adsorbate and (c) evaporation and condensation can occur only from or on exposed surfaces. A detailed balancing of the forward and reverse rates of adsorption on a surface where the number of adsorbed layers is restricted to n by walls of pores having finite diameters leads to the equation: 1 _ vm C X . 1 - (n+1) Xn + n Xn+ [1 2} 1 ‘ X 1 + (C-l) x - c x“+1 where X is the relative vapor pressure, P/Po; v is total volume adsorbed at the measured pressure, P; vm is the volume adsorbed in the monolayer; and C is a constant related exponentially to the heats of adsorption and liquifaction of the adsorbate. Relation 15 [1.2] is the so-called B-E-T 3-parameters equation. It reduces to the Langmuir relationship when n f 1. Under conditions in which the surface is free and adsorption is not limited [n = m], it reduces to the more familiar 2-parameters equation: X 1 + (C-l! X [1.3] v(1-X) : v C v C m m I from which values of vm and C are obtainable from straight line plots of adsorption data in the low-pressure region. In a subse- quent extension of the theory by Brunauer, Deming, Deming and Teller (1940), a much more elaborate equation was obtained to account for isotherms of Types IV and V. The multi-layer theory thus became the first unification of physical—adsorption concepts which applied to the complete isotherm from the monomolecular region through the multi-layer and capillary condensation regions to the saturation pressure. The B-E-T relationship works very satisfactorily in the relative vapor pressure region from 0.05 to 0.3 (Adamson, 1967), and is extensively used as a general method for the determination of surface areas of adsorbents. The considerable success of the B-E-T equation stimulated investigators to consider modifications that would give a better fit to type II isotherms. Among these modifications are those of Pickett (1945), Huttig (1948), McMillan and Teller (1951), Clampitt and German (1958) and Dellyes (1963). A detailed summary of these modifications is contained in a book by Young and Crowell (1962). 16 In spite of its considerable success, the assumptions of the B-E-T theory are not realistic. It does not give the correct variation of either the heat or entropy of adsorption with the amount of gas adsorbed. While the B-E-T assumption correctly reflects the approach of the adsorbed film to bulk liquid prOperties as P approaches Po’ its assumption of localized multilayers is not con- sistent, leading to erroneous configurational entrepy values (Hill, 1960). Related to this is a catastrOphe that Cassel (1944) has pointed out, namely, that the B—E-T model predicts infinite adsorp- tion at saturation. In view of the fact that these same general criticisms apply to the Huttig equation and indeed to the various modifications of / the B-E-T equation, the Opinion of Gorter and Frederikse (1949) that "the kinetical B-E-T theory gives a simple and valuable first pic- ture of the phenomenon of adsorption, but it seems difficult to correct its obvious short comings without destroying the simplicity which perhaps constitutes its chief attraction", seems very appro- priate. Halsey (1950) concludes that the B-E-T and related iso- therms should be regarded as merely convenient algebraic tools for locating the point "B", which, he feels, marks the monolayer stage, Since it is the point of most rapid change in the affinity of the solid for the adsorbate. The Potential Model: At approximately the same time that Langmuir develOped his monomolecular theory, Polanyi (1914, 1916, 1920) formu- lated an entirely different concept known as the potential theory. 17 This concept recognizes the existence of multi-layer adsorption, and considers that there is a pontential field at the surface of a solid into which adsorbate molecules fall. The adsorbed layer thus resembles the atmosphere of a planet, being most com- pressed at the surface of the solid and decreases in density outward. Polanyi defined the adsorption potential at a point on the solid, as the work done by adsorption forces in bringing an adsor- bate molecule from the vapor phase to that point. This work is conceived as a work of compression, and mathematically is given by the so-called hydrostatic equation: 6 = I V dp [1-4] where e is the adsorption potential at a point where the density of the adsorbed film is p; p0 is the density of the gas phase; and V is the molar volume of the adsorbate. It is fundamental to Polanyi's theory that the adsorption potential at any given point is charac- teristic of the adsorbent alone and is temperature independent as well as unaffected by the presence of foreign molecules. It is hence possible to map out the entire adsorption space into a number of equipotential surfaces, the nature of which would be character- istic of the adsorbent. On qualitative grounds, the potential theory appears to be fundamentally correct. It accounts for the empirical fact that systems at the same values of RgT 1n (P/Po) are in essentially corresponding states (Brunauer, 1945) and that the multi-layer approaches the bulk liquid in prOperties as P approaches P0. 18 However, in specific treatments, functions must be assumed to represent the equation of state in the adsorbed phase. Since the adsorbent - adsorbate complex still remains a major thermo- dynamic puzzle, functions used can only be approximate. From this point of view, the original formalism of Polanyi its subse- quent modifications by Frenkel (1946), Halsey (1948), McMillan and Teller (1951) and Hill (1952) together with the adjoint so-called two-dimensional film theory of Harkins-Jura (1944), must be regarded as still somewhat primitive in Specific applications. Capillary Condensation: It has long been recognized that the vapor pressure over the miniscus of a liquid contained in a narrow capil- lary is lower than the vapor pressure of the free liquid at the same temperature, provided that the liquid wets the capillary and forms a concave meniscus. In other words, a vapor is liable to condense in a capillary at a lower pressure than it would on a plane surface. This phenomenon is called capillary condensation. The quantitative relationship between the vapor pressure, P, over a liquid confined in a capillary and the corresponding satura- tion vapor pressure in a free space at the same temperature was given by Lord Kelvin (1871) in the form: 2 O V cos a) [1 5] r _ 1n (P/PO) = where o is the surface tension of the liquid; V is the molecular volume of the adsorbate in its liquid state; r is the radius of the cylindrical capillary; Rg is the universal gas constant; T is tem- perature in degrees Kelvin and w is the contact angle. 19 Zsigmondy (1911) and later Foster (1932) applied the capillary condensation theory to relationships between adsorp- tion and pore structure in porous adsorbents. They argued that in porous structures, the same relationship between vapor pressure and meniscus radius exists as in the case of ordinary cylindrical capillaries. As the equilibrium pressure is increased in an adsor- tion experiment, condensation occurs in successively larger pores. This rationalization leads directly to the calculation of pore-size from adsorption isotherm data employing the Kelvin equation [1.5]. The principal assumption is that the adsorbate exists as a con- densed liquid in the pores of the adsorbent and has preperties characteristic of the bulk liquid phase. Zsigmondy (1911) attributed the hysteresis phenomenon characteristic of adsorption in porous systems, to contact angle hysteresis due to impurities. While this may well be true in some cases, it fails to account for systems having retraceable closed hysteresis 100ps. The schemes of Barret et a1. (1951), Pierce (1953) and Cranston and Inkley (1957) for calculating pore-size distributions from adsorption isotherms represent refinements of Zsigmondy's original theory. In Spite of the warning by Everett (1958) that the bundle of capillaries model can be outrageously wrong for real porous systems, these methods constitute in many cases, the only available estimation of the real pore structure. The "ink bottle" concepts of hysteresis as advanced by Kraemer (1931), McBain (1935) and Cohan (1938, 1944) remain to date, the most plausible qualitative description of the phenomenon. These too, are founded on the capillary condensation theory of Zsigmondy. 20 In conclusion, while the applicability of the Kelvin equation and the assumed bulk liquid prOpertieS of the adsorbed film remain subjects of controversy, most investigators (see Brunauer 1945, Adamson 1967) seem to agree that capillary condensa- tion plays some role in physical adsorption. In general, it is believed to come into play approximately in the region of hysteresis. The Polarization Theory: In the interest of completeness, the Polarization theory of de Boer and Zwikker (1929), and Bradley (1936) is briefly taken up here. This theory explains the adsorption of non-polar molecules on ionic adsorbents by assuming that the upper- most layer of the adsorbent induces dipoles in the first layer of the adsorbed molecules, which in turn induce dipoles in the next layer, and so on until several layers are built op. Brunauer (1945) severely criticized the polarization theory on the ground that the effect was not large enough and consequently the theory has largely been ignored. Even though some recent work by Bewig and Zisman (1964) and Benson and King (1965) suggest that this neglect may be. either mistaken or premature, Hill (1952) points out that there exists no satisfactory theory of the liquid state, even for monatomic liquids, and the detailed treatment of a liquid in a combination electrical and dispersion long-range force field is still far away. SecondaryfiTheories The models to be taken up in this section of the review, are designated secondary in so far as they are in varying degrees sub- ordinate to one or in some cases a combination of the primary models 21 previously discussed. In the context of the present interest, these theories are especially relevant because each one of them has focussed on water vapor as the adsorbate and either polymeric or organic solids as the adsorbent. Although as early as 1882 Mfiller had preposed an equation to predict the adsorption of water vapor by textile fibres, his equation turned out to be quite valueless (Swan and Urquhart, 1927) because his arbitrary assumptions - (a) a linear relationship between the amount of vapor adsorbed and relative vapor pressure and (b) zero adsorption at boiling point of water - were unsound. The Smith Equation: It was not until 1947 that a partially suc- cessful treatment of water sorption by polymers was formulated by Smith who recognized the existence of two principal classes of sorbed moisture: (a) bound moisture, (ab), held on the adsorbent surface by intermolecular forces in excess of forces responsible for condensation and (b) normally condensed moisture, (ac). He assumed that the relationship between ab and P/Po can be approxi- mated by the Langmuir equation. While accepting the multi-layer concept of Brunauer, Emmett and Teller as the structural framework of ac and further adepting the B-E-T position that the multi-layer is essentially in the same thermodynamic state as the bulk liquid, he derived the following expression for ac: ac = - a’ 1n (l-P/PO) [1.6] and summed ab and ac to obtain: _ = u ’ - a - ab + ac ab a 1n (1 P/PO) [1.71 22 where a, is defined as the adsorbed mass per gram of the adsorbent in a unimolecular layer of normally condensed moisture, and a is the total specific adsorbed mass. In order to arrive at equation [1.6], Smith assumed that at any stage of sorption, the completed fraction of a, is a measure of the total potential evaporating surface - a perfectly valid assumption. However, his subsequent assumption that this fraction, (9), is equal to P/Po is thermo- dynamically unsound. While it is true that in the assumed thermo- dynamic state of ac, both 9 and P/Po represent some measure of the escaping tendency of the vapor molecules, it can be shown that they do not reduce to an identity. All that can be deduced from thermo- dynamics is that: 9 = 8(P/PO). Invoking (a) Lewis and Randall's representation (Lewis and Randall, 1961, p. 147) of fugacity as the molal free energy and (b) Cibb's definition of the surface free energy as the work done against surface tension forces by the Spreading two - dimensional pressure on the adsorbing surface (Gregg and Sing, 1967, p. 234), the following approximation can be written: n 9 = AF = -R T ln (P/P ) [1.8j g o where 9 is taken as a measure of the surface on which the spreading pressure, (H), is active. Equation [1.8] suggests that Smith's fundamental simplification cannot be thermodynamically justified. It must therefore be concluded that even though the Smith equation has been found in certain cases (for example Becker and Sallans, 1955) to fit experimental isotherm data remarkably well in the 23 region 0.5 S P/Po S 0.95, his exponential representation of the multi-layer region is basically empirical. That empirical func- tions can be selected to fit a wide variety of isotherms has been demonstrated by Strohman et al. (1967) and Do Sup Chung et a1. (1967), both of these works will be elaborated upon later in this section. Other weaknesseS'of the Smith equation include (a) its inability to give the correct variation of either the heat or entrepy of adsorption, (b) its inability to account for hysteresis and (c) the fact that temperature dependence is not inherent in the formulation. The Henderson Equilibrium Equation: Perhaps the best known and most widely used equation for predicting the equilibrium moisture content of biological materials is the semi-empirical equation of Henderson (1952). Using the Gibb's adsorption equation as a starting point, Henderson derived an isotherm equation which can account for the temperature dependence of the experimental curve. His equation is of the form: kTMan (1 - P/P.) = e' [1.9] where T is absolute temperature in degrees Rankine; and k and H are empirical constants. When apprOpriate values of the parameters, k and H, are available, the Henderson equation or its modifications by Day and Nelson (1965) and Thompson et a1. (1967) have been found to fit isotherm data for cereal grains fairly well (see Rodrigues- Arias, 1956). However, the Henderson equation has been found to be 24 quite inadequate for certain biological products (see for example Pichler 1957, Bakker-Arkema 1961, Day and Nelson 1965). One deficiency of the Henderson equation is that it is totally based on thermodynamics, it is therefore not founded on a model and gives no information about the nature of the adsorbent or its surface. The Young-Nelson Hysteresis Equation A creditable effort to construct a theory of adsorption for biological materials to reflect their basic cellular nature was made by Young and Nelson (1967). These investigators considered the cell as the ultimate basis of adsorption and recognized the existence of three modes of adsorbed moisture: (a) the unimole- cular-bound moisture of Langmuir, (b) the normally condensed or multiple-layer moisture of Brunauer, Emmett and Teller; and (c) an adsorbed moisture which results from a diffusional passage of moisture into the inner cell and which on account of the irreversi- bility of the diffusion process is responsible for the occurence of hysteresis. The development of a representative expression for the first category of moisture is kinetical; being in essence identical to the B-E-T process. Their develOpment of an expression for the normally condensed moisture, with only slight differences, parallels the Smith method of attack and implies the same basic assumptions. It is therefore not surprising that they arrived at an equation which has the prOperties of a combined B-E-T-Smith equation (Strohman et al., 1967) including the advantages of both but impeded by the fundamental empiricism of the Smith equation. 25 Although these authors by their perceptive introduction of the "absorbed moisture" have deve10ped a comparatively straight forward quantitative representation of hysteresis, the simplifi- cations and reasoning leading to their explicit expression for the absorbed moisture effectively destroyed their model. If the ulti- mate cell of a biological product is taken as the basis of sorption, it appears logical that moisture transfer across the semi-permeable cellular wall can take place only as a result of osmosis. From thermodynamic considerations, the osmotic pressure necessary to in- duce such transfer is given by: (see Guggenheim, 1967, pp. 183-184): RT . __a_ H _ V1 1n (Pl/P) [1.101 where H is osmotic pressure in atmospheres, Rg (the Universal gas constant) = 82.06 cc.atm. °K-1mole-1; V1 is the volume of adsorbed moisture in cc.; and T is absolute temperature in degrees Kelvin. For the 30°C isotherm, the values of H correSponding to the relative vapor pressure values of 0.1, 0.3, 0.8 and 1.0 are respectively: 1.6 x 106, 4.6 x 105, 3.46 x 104 and 0.0 atmospheres. The values of V1 corresponding to the above P/Po values are obtained by conversion of the adsorption data for corn. The conversion constants will be developed later in this work. The specific values of V1 used are 0.0351, 0.0654 and 0.1590 cc. Since the osmotic pressure must be exceeded or at least balanced for transfer across the cell wall to take place, it appears clear that except at saturation, infinite layers of adsorbed molecules will be necessary to create a suffi- ciently large hydrostatic force. Moisture transfer across the membrane 26 cannot therefore be justified prior to saturation. On the other hand, with a predicted zero osmotic pressure at saturation, all the normally condensed moisture on the surface would be expected to diffuse into the cell. In summary, our present line of reasoning allows for no adsorbed moisture except at saturation, at which point the normally condensed moisture begins to diffuse into the cell. This moisture passage should continue until the cell fills to its maximum absorptive capacity. The Young-Nelson picture of the absorbed moisture does not therefore agree well with thermodynamic reasoning. It must, however, be stressed that in spite of its apparent inability to stand up to its theoretical implications, the Young-Nelson model of hysteresis represents the best quantitative effort yet to solve explicitly the problem of hysteresis in water adsorption by biological materials. The ChungePfost Eqpation: The general framework of the potential theory was utilized by Chung and Pfost (1967) to develop an iso- therm equation for cereal grains and their derivatives. The equation is of the form: _'_‘.‘__ - 1n (P/Po) — RgT exp( BM) [1.11] where M is the specific adsorbed mass and the parameters A and B are product and temperature dependent empirical constants. Like the earlier but algebraically identical relation of Dubinin and co- workers (1955, 1965), the Chung and Pfost equation is semi-empirical being based on an over-simplified equation of state for the adsorbed film. While specific instances of success in the use of such equa- tions can be cited, their wider usage cannot be justified on theoretical grounds. 27 The Empirical Eguation of Strohman and Yoerger: Strohman and Yoerger (1967) using the Othmer linear plots as a starting point developed a purely empirical isotherm equation for corn. Since they had no working model, their method must be considered as essentially a curve fitting scheme which offers practically no descriptive picture of the phenomenon. While such schemes may become useful in secondary applications such as drying for example, they throw no light on the basic process of adsorption. 1.7 Statement of the Problem While the three basic theories of adsorption and their subordinate equations have failed in their individual capacities to produce a generally satisfactory mathematical formulation of the sorptive response of biological materials when exposed to a vaporous atmosphere of water, their apparent complementary character indicates, as has been suggested by KUhn (1964), that the fundamental concepts may be successfully combined into a unified model of adsorption. _As a broad statement of the problem, it is attempted to con— struct an isotherm equation for biological materials based primarily on the B-E—T and capillary condensation theories and somewhat in- directly on the potential theory. This integrated theory will be justified if the porous nature of a given biological material can be considered fundamental to its sorptive behavior. In the light of the best available knowledge in the area of sorption, it is reasonable to consider the adsorption process by biological products as resulting from mono and multi-layer 28 phenomena up to the incidence of hysteresis; the subsequent isotherm progress into the region of hysteresis is attributed to both multi-molecular adsorption as well as capillary con- densation on the porous solid. 1.8 Objectives The study reported in this dissertation was undertaken for the following specific objectives: 1. To construct, utilizing the B-E-T, and capillary condensation theories together with a working model of their pore structure, a moisture adsorption model for biological materials. 2. To derive, using the general framework of the potential theory in conjunction with the fundamental laws of thermodynamics, an explicit density function for adsorbed moisture. This function is then integrated into the moisture adsorption model to obtain a generalized iso- therm equation for porous biological materials. 3. To verify the validity of the prOposed model by showing that: (i) The heat of adsorption and its variation with the amount of vapor adsorbed are inherent in the formulation. (ii) Sorption hysteresi§_is both qualitatively and quantitatively expressible within its governing framework; (iii) The derived isotherm equation is experimentally correct. 29 II. THEORETICAL CONSIDERATIONS 2.1 Introduction In this chapter the framework of a generalized model of adsorption is formed. The B-E-T and Capillary Condensation theories are combined into an integral isotherm equation accounting for mole— cular and capillary adsorbed moisture in porous biological materials. In order to solve the integral equation explicitly, it became neces- sary to: (a) characterize the pore structure of adsorbing biological products mathematically; (b) define an explicit density function to satisfy the fundamental postulates of the adsorption potential theory as advanced by Polanyi. As a necessary self-consistency test for this line of attack, a con— cept of sorption hysteresis is constructed within the generalized framework, thereby making possible the quantitative estimation of hysteresis in swelling biological materials. 2.2 A Physical Model of Water Adsorption by Bio-materials The physical picture advanced here is illustrated in Figure 2.1 which shows an organic tissue in a vaporous atmosphere of, for instance, water at a specified pressure and temperature. The tissue is visualized as a random network of small pores, the diameter of one of which has been magnified for illustrative purposes. The pores are interstices between ill-fitting cellular building blocks which make up the tissue. Although these intercellular passageways certainly are not cylindrical, a useful picture can be made in which they are 30 07:301W .>Efiumuuwsfiflw new pmwmwcwwe ma open one we umuoEme wLH .mouoa hogawwuuw .Hfimem mo xuosuoc Eopcmu m on msmmflu omcmwuo no mo cofiumucmmmuaou HMHHOuoLm .H.N muowwm MCDFQOZ ommmomod 62.30...» $144453 KSJNO_2thhZ. omr=20<2 mm><4 1:332 ommmomo< mo mmmzxofk « manic V mo z.>.._mv. L 31 Jim‘s... hi“ ,. Um, . I wflé treated as cylinders, parallelepipes (slitted pores) or inter- connected spheroidal "ink bottles", with rough walls and inter- secting with other pores. The sorption of water vapor by the tissue can now be thought of as a process by which the pores are isothermally filled or emptied of water vapor under the influence of surface forces active on the pore walls. While this simplified picture classifies the tissue as essentially a porous body, it does not preclude its basic organic and biological nature. The position is taken here that unique modes of behavior such as respiration, and ”water-active” sites which characterize biological materials in their natural state, exert a tremendous influence on the nature and distribution of surface forces. In consequence, it is logical to assume that at least some Of the "live processes" find commensurate expression in the energy and entrOpy configurations associated with the sorptive process. The point is stressed here that in the gross treatment of physisorption in biological materials, the separate and detailed consideration of associated life phenomena may become superfluous since such processes are already inherent in energy or entrOpy terms. As a micrOporous body, the pore filling process proceeds due to the Operation of two mechanisms: (a) molecular adsorption, and (b) capillary condensation. At any stage in the process, its adsorp- tive capacity is measured by the adsorptively utilizable volume of the constituent pores. The practical objective Of the physical model being developed is to characterize these pores sufficiently in terms of size, and size distribution to make the computation of the afore- mentioned volume possible. 32 2.3 Generalization of Molecular and Qgpillaryyédsorbed Moisture 2.3 a. Pore Characterization — defining the size f.g pore filled by Molecular and Capillary Adsorption: In its strictest sense, the term "capillary condensation" is applied to the particular adsorption mechanism described by the Kelvin equation. In this mechanism the equilibrium vapor pressure of a liquid in a cylindrical capillary is reduced below its saturation value by negative hydrostatic pressure arising from tensile components of curved surfaces of tension of the liquid, i.e. from the menisci. The reduc- tion in pressure is related to the radius of curvature of the menisci by the Kelvin equation (1.5). If complete wetting is assumed to occur so that q>is equal to zero, equation (1.5) can be reduced to the simpler form: _ 2 0'6 rk ‘ RgT 1n(Po/P) [2°3’1] where rk is the radius of curvature of a hemispherical meniscus. Although eQUation (2.3.1) is basically descriptive of such microsCOpic phenomena as capillary rise and depression, Zsigmondy (1911) and later Foster (1932) applied it to relationships between adsorption and pore size in micrOporous adsorbents. Their principal assumption was that the adsorbate behaved as a single-component liquid in a poten- tial field inside the micrOporous capillary of the adsorbent and pos- sessed prOperties characteristic of the bulk liquid phase. In view of the undefined thermodynamic state of the adsorbent-adsorbate complex, this rationalization remains open to question. However, investigators consider that the Kelvin equation can be used with reliability for the 33 calculation of pore size along the desorption path of the isotherm (Flood, 1967, p. 55). This is particularly true in the case Of H20 where adsorbent contraction accompanying adsorption hysteresis is indicative of (a) the presence of a hemispherical meniscus, and (b) the active Operation of a negative hydrostic pressure - both of which are consistent with the Kelvin mechanism. Wheeler (1945) preposed an improved theory which took into account the effects of multi-layer adsorption. Two new assumptions were made. First, it is assumed that at any point on the desorption branch of the isotherm, all pores larger than a certain radius R, are covered with an adsorbed multi-layer of thickness t, whereas all pores smaller than R are filled by capillary condensation. Secondly, since all unfilled pore walls have an adsorbed layer of thickness t, the radius of the meniscus in a filled pore, where it joins a larger pore, is assumed not to be the pore radius R, but a smaller radius (R-t). This means in effect that under conditions of capil- lary adsorption it is not the pore of the true physical radius R which is important, but rather, a pore whose radius has been effec- tively diminished by the thickness of an adsorbed multi-layer. Wheeler thus argued that the Kelvin equation applied not to the actual pore radius but to the effective radius Of the "inside annular tube” left after multi-layer adsorption has taken place. The maximum true pore radius R, which will be filled by both molecular adsorption and capillary condensation at a relative vapor pressure P/PO is thus given by: = 2.3.2 R t + rk 1 1 34 where t the thickness of adsorbed multi-layer, yet to be defined. r the Kelvin radius defined by equation (2.3.1). The k summed radius is shown diagrammatically in Figure (2.2). Because of its Kelvin component, the Wheeler equation (2.3.2) is applicable only to the desorption branch of the isotherm. How- ever, Cohan (1938) and Coelingh (1938), working independently, con- cluded that capillary condensation occurred along both branches of the isotherm 100p, and explained hysteresis in terms Of the different shape of the meniscus during adsorption and desorption. Along the desorption branch the Kelvin mechanism is assumed - the meniscus being hemispherical - but along the adsorption branch, the meniscus assumes a cylindrical shape, the pore being Open at both ends (see Figure 2.3). Cohan by (a) specializing the Young-Laplace equation for the case of a cylindrical meniscus (see Gregg and Sing, 1967, p. 149), and (b) employing the same basic thermodynamic argument as Kelvin, arrived at the equation: (IV C = R T 1n(P /P) [2.3.3] g 0 where rC = the radius of curvature of a cylindrical meniscus (the Cohan radius). The Wheeler equation (2.3.2) can be now generalized into the form: R = r + t [2.3.4] where r = rc for adsorption; and r = rk for desorption. Having already defined r in terms of the variable, P/PO, the thickness t needs to be similarly defined in order to complete the characteriza— tion of R. 35 Figure 2.2. Figure 2.3. Capillary menisci: or Kelvin equation. :\ 36 1 I’III’I’I”III77III’IIIrf77r7'II'll/rilllrrifiirr i Cylindrical capillary of radius R, showing the "inside annular tube" r which may be given either by the Cohan \1 (a) Hemispherical (Kelvin) meniscus. (b) Cylindrical (Cohan) meniscus. -*- 2.3 b. The thickness pf the adsorbed multi-layer: Halsey (1948) employed the concept of COOperative adsorption to refine the B-E-T theory. This concept implies that adsorbate molecules influence each other by horizontal interaction during the adsorption process. He derived the following semi—empirical relation for the adsorbed multi-layer thickness: 2 Qst S RgT 1n(Po/P) ] cw) = T [ [2.3.5] where T - average thickness of a single layer of adsorbed molecules Q = the isosteric heat of adsorption Q = an empirical constant dependent on the adsorbate, the adsorbent and possibly temperature. The most direct empirical way of estimating t, is to determine by the standard B-E-T procedure the monolayer capacity Xm of the adsorbate. The adsorption X at any pressure can then be converted into the thickness of the film, by employing the relation. cw) = (i—‘-) . 'r [2.3.6] In or t(A°) = n T [2.3.7] where n is the number of molecular layers adsorbed. It needs to be emphasized that the value of T is not necessarily the same as the molecular diameter, d of the molecules. The relationship between the two depends on (a) the mode of stacking of successive molecular layers in the adsorbed film as illustrated by the variable configurations postulated in Figure 2.4, and (b) the degree of adsorbate compression 37 0969 s‘s‘e‘s‘o’e’e’4‘4 O. 3.0.0.0331... (a ) Figure 2.4. Packing of adsorbed molecules in successive molecular layers, to illustrate variable stacking configurations. 38 within the adsorbed film. Two different values for nitrogen of 3.5(A°) (prOposed by Lippens, Linsen and deBoer, 1964) and 4.3(A°) (prOposed by Shull, 1948) are clearly illustrative of the state of controversy existing on the subject. For the adsorbed water mole- cule, an even greater degree of uncertainty must be entertained with regard to the value of T because of the high degree of compres- sion associated with water adsorption (see Babbitt, 1942; Stamm, 1938). In Figure 2.5, X/Xm is plotted against P/Po for a number of biological products. Superposed on this plot is a curve described by the modified Halsey equation: 1 75 I = ° 2. . mm [W3 E 383 Qst where the quantity (E1?) in the Halsey relationship is approximated 8 by the constant 1.75. This theoretical curve fits the empirical X/Xm values fairly well. The individual points are however, scattered around the composite curve. This means that some amount of uncer- tainty exists in the value of X/Xm defined by equation (2.3.8) for a given sample. If this degree of uncertainty is considered within limits of acceptable error, then t is defined by the Halsey equation in the form: 1.75 t(A°) = 3.2[31—12713] [2.3.9] where for the water molecule, T is taken to be approximately equal to d for lack of a better approximation. 39 g THEORETICAL CURVE (eqn- 2'3-8) J L 1 J 1 J 1 L a L + 0 0.0 0.: 0.2 0.3 0.4 0.5 0.6 0.7 0.8 09 LG P/P. Figure 2.5. Plot of )(/)(m against P/I’O. a , wood 20°C isotherm. A, cotton 10°C isotherm. 0 , cotton 20°C isotherm. A, cotton 30°C isotherm. X, corn 4°C isotherm. [3] , Ground freeze-dried beef 10°C isotherm. I . ound :meze-driml hoof 22.2”": isotherm. 40 Equation (2.3.9) in a slightly different form was used by Shull (1948), Wheeler (1955), Dollimore and Heal (1964), and Viswanathan and Sastri (1967) for the calculation of the adsorbed multi-layer. It is adapted in the present study in preference to the B-E-T 3-parameters equation from which n = X/Xm is derivable, for a number of reasons. It is consistent with the concept of mobile adsorbed film. It allows for the variation of heats of adsorbtion with amount adsorbed. It has a form convenient for usage in a larger context. 2.3 c. The Isotherm Equation: With R fully defined as a function of relative vapor pressure, the equation for the adsorption isotherm is deve10ped as follows: The Wheeler equation (2.3.4) specifies the relative vapor pressure at which condensation will occur in a capillary of a given size. Such condensation will also occur on the adsorbed layers of water molecules already covering the concave surfaces of a porous system. If such condensation leads to a significant reduction in the radius of curvature of a concave surface or Open ended capillary, condensation will continue on this surface as long as the radius of curvature continues to reduce. In this discussion, a pore will be defined as any void region in a porous material which is partially or completely filled by the combined consequence of molecular and capillary adsorbed moisture. Also the pore radius will be considered to mean the radius of curvature of the surface on which the above described filling-process started to occur. This radius is defined by the generalized Wheeler equation (2.3.4). 41 If a porous system of organic tissue has a given distribu- tion or pore radii, one can say that at a specified vapor pressure, pores of radius equal to or less than the value determined by the Wheeler equation (2.3.4) will be completely filled. That is, if the volume of a pore in a porous system is expressed by the function vR = Cpl(R) [2.3.10] and if the number of pores of radii between R and R + dr is given by dN = m2(R)dR [2.3.11] then, the volumes of all pores with radii between R and R + dR is: d vR = deN = m1(R) x m2(R) dR [2.3.12] Thus, the sum of the volumes of all pores whose radii are equal to or less than R is given by the integral: R Va = JR“. [ 01m X 202m) 1 cm [2.3.13J where Rm, the lower limit of integration is used in recognition of the so-called "molecular sieves" effect which assumes that adsorbed molecules in the first approximation, cannot penetrate into pores smaller than their own diameter, 2Rm. Equation (2.3.13) is essentially an integral isotherm equation defining the adsorptive volume of the porous adsorbent. In order to solve this integral equation explicitly, it is necessary to obtain explicit analytic functions for (a) the pore geometry, w1(R) and (b) the pore size distribution, w2(R). 42 de Boer (1958) has shown that a wide variety of geometric models are possible. However, only two broad classes of these geometric shapes are consistent with the assumptions of the present deve10pment. These geometric forms represented diagrammatically in Figure 2.6 are: (i) The Cylindrical type pore model for which: 1T L(R) R2 [2.3.14] c0101) where L(R) variable length of pore. (ii) The Interconnected Spheroidal "Ink Bottle" pore model for which: cp1(R) = 5:;- rr R3 [2.3.15] Wheeler (1945) suggested and Shull (1948) demonstrated that pore size distributions may be represented by simple analytical relationships of: (i) The Maxwellian type for which: .3 @2(R) = A exp [-R/Ro] W 0 (ii) The Gaussian type for which: 2 w2(R) = A exp [- 82( - 1) J wlw o where A, B and R0 are constants. Gregg and Sing (1967) have further suggested the possible utilization of: (iii) The Log:Normal distribution for which: m2(R) = A exp [- 82(1n R/RO-1)2] 43 (a) g (b) Figure 2.6. Shapes of capillaries. (a) Cylindrical capillary (b) Interconnected spheroidal "Ink Bottles". 44 No closed form solution to equation (2.3.13) is possible when any one of the above analytical functions is used to define m2(R), the worst of them is mathematically intractable, the best leads to unwieldy end results. In the attempt to overcome this difficulty, Foster (1948) and Kuhn (1964) recommended a power function distribution of pore radii for which: cp2(R) = KlRY [2.3.10] where the symbol y is an exponent dependent on the product. Equa- tion (2.3.16) is essentially of an exponential character, and for the most simple cases, it reduces to a simple power function in which y can assume constant values, being positive or negative, integral or fractional. A detailed computation of the pore-size distribution of several biological products was undertaken using a modified Cranston and Inkley (1957) scheme. The results - to be presented in chapter 4 - show that if the spheroidal "ink bottle” geometric model is assumed, m2(R) [as defined in equation (2.3.11)] is well described by the power law equation (2.3.16). In view of its central impor- tance to the deve10pment presented, the above mentioned computational scheme together with the results need to be considered here in detail. However, in order to maintain continuity in the line of thought, it is desirable to postpone this side detail for consideration in a later section. Combining the "ink bottle” pore geometry with the power func- tion distribution of pore radii in equation (2.3.13) yields: 45 R v = -% U K I R3+Y dR [2.3.17] Equation (2.3.17) integrates to: = ,3 n_ n va n ( R Rm ) [2.3.18] where 5:5-TTK [2319] .3 3 1 o. and n = 4 + Y [2.3.20] Equation (2.3.18) is a simplified geometric expression of the adsorptively utilizable volume of a porous adsorbent. The exponent n can have positive or negative, integral or fractional values. Substitution of the generalized Wheeler equation (2.3.4) into equation (2.3.18) and using the Cohan equation (2.3.3) together with the specialized Halsey equation (2.3.9) yields: l, _ _ E 1.75 3 0 v n Va — n { [3'2 (1n Po/P) + RgT 1n(Po/Pm)J L -— n - 1.75 7 0 V 1 “ L3°2 (In P /P) + R T 1n(P /P ):J } [2°3°21J o g o m where Pm is the vapor pressure corresponding to the adsorbed mono- layer or point "B" of the isotherm. Equation (2.3.21) is an isotherm equation relating the specific adsorbed volume of the adsorbent, Va’ with relative vapor pressure. It shows strange deviations from experience in that it gives zero adsorption, Va = O at P = Pm, and negative values of Va at pressures, P < P5. This means that isotherms calculated with equation (2.3.21) 46 must start not at P/Po = 0, but at a low relative vapor pressure Pm/P0 > 0. This irregularity stems from the fact that the choice of the lower limit of integration, Rm, in equation (2.3.13) while conforming with the specifications of the "molecular sieves" effect, overlooks the contribution to Va due to the partial filling of the first mono-layer in the vapor pressure range 0 S P S Pm' Yet, the choice of Rm is substantiated in the now accepted argument (Wheeler, 1955) that the Kelvin or Cohan radius is inapplicable in the low pressure range 0 S P < Pm. In order to correct for this irregularity, a shift of the reference axes of the isotherm plot was performed, thereby reducing equation (2.3.21) to the form: =5 n_ n Va n (Z X ) [2.3.22] where g _ 1 75 0 V Z = 3.2 (—L———) -————————- [2.3.23] 1n X1 RgT ln x1 )(=32(.1_'7_§_)%+_G_V__ [2324] ° ln x2 RgT ln X2 ° ' X1 = (P() + Pm)/(P + Pm) [2.3.25] x2 = (P0 + Pm)/Pm [2.3.20] Equation (2.3.22) is a volumetric isotherm equation defining the specific adsorbed volume due to molecular and capillary phenomena within the intermicellular capillaries of the tissue. It is reduceable to its gravimetric equivalent by a process which converts the specific adsorbed volume to specific adsorbed mass. This can be accomplished by introducing the apprOpriate density term into equation (2.3.22) 47 to obtain the terminal relation: Ma . 0% (2” - A") [2.3.26] where, in view of the demonstrated variable compressibility of the adsorbed water "film" (Katz, 1933; Stamm and Seborg, 1935; and Stamm, 1938), p is an undefined function which is dependent on the magnitude of the intermolecular forces, the surface-impressed pressure and possibly the isotherm temperature. Polanyi's Adsorption Potential theory describes the adsorbed multi-layer as resembling the atmos- phere of a planet with the highest compression at the surface of the solid and the density falling off outwards (see Figure 2.7). Babbitt (1942) insisted that any theory of adsorption must account for this adsorption compression. Stamm and Seborg (1935) demonstrated that actual adsorption compression values can be obtained, and that adsorption compression extends to the fiber saturation point. A den- sity function for the adsorbed water film reflecting these points of view needs now be formulated to augment the derived isotherm relation (2.3.26). 2.4 Adsorption Compression on Adsorbing Bio-Materials Even before Polanyi (1914) first expounded his now famous theory of isopotential contours within the adsorbed film, a large number of investigators had preoccupied themselves with detailed theoretical and experimental study of the adsorbed phase. Among sudh early investigators are Rose (1849), Jungh (1865) and Parks (1902) who speculated that adsorbed water molecules on the surface of a 48 / ___ _~ / —-—— "" "‘ —' ABSORPTION / ’ ‘ \ ./ SPACE / ”—— ‘g’ ‘—' .-—_ \ un— / / / // ,_ \ // (ttWWW Hm ADSORBENT Figure 2.7. Polanyi's isopotential contours. The space between each pair of equipotential surfaces corresponds to a different degree of adsorbate compression. 49 solid have undergone a definite thermodynamic change in state and are held against the surface by a force which is expressible in terms of pressure and compressibility. Calorimetric measurements of the heats of adsorption undertaken by Lamb and Coolidge (1920), Patrick and Grimm (1921) and later supported by the extensive thermodynamic calculations of Harkins and Ewing (1921), estimated values of the compressive force in the range of 37,000 to 100,000 atmospheres. From specific volume determinations for cotton in different liquid and gaseous media, Davidson (1927) calculated values of the average compression of sorbed water of approximately 2000 atmospheres on the basis that compression takes place entirely in the water and that the adsorbent is essentially incompressible. Katz (1933) has shown that a qualitative pr0portionality exists between the heat of swelling of binary aqueous systems and the corresponding volume change. Gibson (1934) gives the relationship between the specific compression of water, the internal or intrinsic pressure change, and the externally applied pressure for aqueous solutions of certain sulfates. Measurements are available giving the contraction in volume as a function of moisture content for cellulosic materials. Thus, Stamm and Hansen (1937) employed the Gibson compressibility relation- ship for water to obtain specific volume contractions for wood, pulp and cotton. Filby and Maass (1932) calculated the apparent density of water adsorbed on cotton cellulose as a function of moisture content. They concluded that below the 3% moisture level an apparent density of 2.6 gm./cc. is obtained. The apparent density value falls off with increasing moisture content, until above the 9% moisture level 50 where the adsorbed water appears to have its normal bulk liquid density. This result was contradicted by the work of Stamm and Seborg (1935) who found that the density of water originally adsorbed is 1.3, and demonstrated that adsorption compression extends to the fiber saturation point. Ewing and Spurway (1930) obtained apparent density values for water adsorbed on silica gel. Their values extend all the way from 1.03 at 1% moisture content to 0.54 at 7% moisture level. Morrison and McIntosh (1945) obtained the apparent density of water adsorbed on four different charcoals using the helium displacement method. Their results which exhibited considerable inconsistency, showed a notable difference in apparent density values of water between the adsorption and desorp- tion branches. In another study, Tuck, McIntosh and Maass (1947) investigating the density of various adsorbates on charcoal, con- cluded that in view of the uncertainties which they observed, the utility of apparent density values of adsorbates in checking con- cepts in physical adsorption is limited. From the foregoing review, it is apparent that while the broad concept of adsorption compression seems to be well established among investigators, considerable uncertainty exists regarding compressibility and apparent density values of adsorbates. This uncertainty must account, at least in part, for the seeming lack of interest on the part of investigators to come up with a descriptive density function for the adsorbed phase. The compressibility rela- tion of Gibson notwithstanding, there exists, to the knowledge of the author, only two such equations. The first is the density 51 relation of Stamm (1938), which expresses adsorbate density in terms of the "true" and "apparent" specific gravities of the sorbent. The second is a computational scheme using the Lowry and Olmsted (1927) modification of Polanyi's theory to generate the so-called characteristic curves from empirical isotherm data (see Brunauer, 1945, pp. 101-104). The character- istic curves are then integrated numerically to obtain the density values. Because of the difficulty of obtaining accurate Specific gravity values for porous substances, the Stamm relation does not lend itself for usage here. The Polanyi characteristic curves scheme on the other hand, is only as good as the simplified equa- tion of state which must be assumed for the adsorbed phase. For their work, Lowry and Olmsted used a modified form of the van der Waals two-dimensional equation of state given by van Laar (1924). Even though Lowry and Olmsted (1927) and De Vries (1935) appear to have successfully used the method to obtain reasonable density values for C0 adsorbed on charcoal, its direct application to 2 H20 systems does not seem to be as yet possible since the van Laar constants for the water molecule have not been determined. Because of the urgent need for an easily applied adsorbate density relation, a derivation of such a function will now be presented. 2.4 a. g Density Function for Sorbed Water: Consider a closed vessel (Figure 2.8) of volume V containing the sorbent and a fixed number of molecules, n, of the sorbate at the temperature T and pressure P. If Vm denotes the molar volume 52 ;\\\\\\x\\\\xx\ x;\\\\\\\\ \\x\' x \\x K \ k \ K N N \ \ x N \ x x x \ T 111111111 I I Ansoasms T SYSTEM p n SIDE PURE n CYLINDER ADSORBATE m \ \\\\ ;\\\\x\\\\\\\;;\\ x\\\\\\\ \\\xx 7r/ir/friI/r/Ir/lr/I/7r4 Figure 2.8. iii/TITIIIII/II (a) (b) Diagram showing two thermodynamically identical systems. (a) Contains an active sorbent (h) Contains the pure sorbate 53 of the gaseous sorbate and the volume of the sorbent is assumed to be negligible, then: a V = (n — n ) Vm [2.4.1] a where n is the adsorbed excess due to sorption over and above the amount of gas which would be contained in the same volume, at the same pressure and temperature, in the absence of the sorbent. From.Maxwe11's relation: (bS/BP)T = -(0V/BT)P [2.4.2] combining equations (2.4.1) with (2.4.2) gives: - (aS/BP)T = (BVIBT)P a a = (n-n )(avm/aT)P-(0n /0T)Pvm [2.4-3] also 8 3? (as/62)., = - (as/an )T(507)T consequently a _ a (as/an )T - (BS/0P)T (BF/5n )T a a a a - (n-n )(BVm/BT)P(aP/dn )T - Vm(dn /dT)P(5P/dn )T a a = - 5 2. . (n n )(de/GT)P( Plan )T + Vm(5P/8T)na [ 4 4] Now compare the system described above with another system (shown in Figure 2.8.b) consisting of a vessel of the same volume V containing the same gas at the same temperature and pressure, but without any sorbent. If primed symbols are adOpted to represent quantities relating to this second system when they may differ from those relating to the first system, then: 54 n, = n - na [2.4.5] where n, denotes the number of molecules of pure gaseous sorbate remaining in the vessel after the amount n3 has been lost iso— thermally through the lateral cylinder. Moreover, since both systems are in equilibrium and the gas is in identical condition, the chemical potentials are the same, i.e. U = u [2.4.6] If the molar entrOpy change in the first system per unit change in na resulting from a decrease in the pressure, is denoted by ASm, then AS = (aS/ana) m T a a = _ 2.4. (n n )(avm/aT)P(6n )T + vm(aP/aT) a [ 7] n If the molar entrOpy change in the second system corresponding to the same decrease in pressure is further denoted by A’Sm, then, I I a . = 2. . ' A Sm n (avm/aT)P(0P/0n )T [ 4 8J Subtracting equation (2.4.8) from equation (2.4.7) and using equation (2.4.5) gives: — i = a — a : [ Asm A Sm (BS/5n )T (as/0n )T vm(0P/0T)na [2.4.9] Multiplying equation (2.4.9) by T gives: T as - T A’ s = T(5P/8T) v [2.4.10] m m nam 55 It follows immediately from equation (2.4.6) that: A u = A’ u [2.4.11] From the fundamental equations of thermodynamics (see Guggenheim, 1967, p. 24), (1 G. .g—i‘. (“a/T) = - -—- .. £5 a: - 357 [2.4.12] T T It follows immediately from equation (2.4.12) that: (1. CI. r A s + A n = A a“ [2.4.13] Subtracting equation (2.4.11) from equation (2.4.10) and using equation (2.4.13) yields: A Hm - A Hm = T(3P/8T)navm [2.4.14] The left hand side of equation (2.4.14) may be taken to represent the equilibrium molar enthalpy (heat) of sorption. It is the heat which must be supplied per unit change in n8 resulting from decrease of pressure under isothermal quasi-equilibrium conditions, minus the heat that must be supplied to the second system when the pressure is identically reduced by the same amount. It needs to be emphasized that every quantity occuring in equation (2.4.14) is experimentally determinable without the use of approximations or extraneous assumptions. Thus the quantity A’sm is a molar work function for a real gas, and AHm and (op/or) a n are obtainable from sorption isosteres. 56 If the sorbed mass in the sorbent-sorbate system is denoted by M; and the corresponding volume by Va then, V a a n vm a (Ma/M) vIn [2.4.15] also p Ma/Va = M/Vm [2.4.16] where p is the density of the sorbate and M.is its gram molecular weight. Multiplying equation (2.4.14) by n8 and using equations (2.4.15) and (2.4.16) gives: 'H a I i- (AHm- 7] 11m) = '1‘(51>/a'r)uava [2.4.17] so that: T(BP/BT)na = -—-—--—-- 2 ° . p (AH-A’H) [ '4 18" Equation (2.4.18) can be expressed in terms of a standard density, p0, such that: o I o BP/bT) a = AH -A H ( n z where the superscript "0" denotes saturation values, and p0 is the density of the sorbate at saturation pressure. The differential terms of equation (2.4.19) can be defined by invoking the Clausius-Clapeyron equation without any loss of accuracy. Thus: (3.; .. 211.123 [2.4.20] na RgT 57 (92) = AH°P0 dTno RgTz combining equations (2.4.20) and (2.4.21) yields: (BF/5T)na = 92. P/P ) (BF/5T)n0 AH°( o Substituting equation (2.4.22) in equation (2.4.19) gives: P AH°-A’H° . Ag 9/0. ' '130‘W) no or 9/90 = u* (AH/AH°) AH°-A’H° *3 where u (P/Po) (Zfi:37§") for pure water vapor A'H° i 0 consequently P AH° “ (————7') Po AH-A H [1* [2.4.21] [2.4.22] [2.4.23] [2.4.24] [2.4.25] [2.4.26] Equation (2.4.24) states that the adsorbate density is a direct function of the isosteric heat of sorption which in turn is a measure of the intermolecular force of attraction. result appears to support the earlier qualitative observation of Katz (1933), that a prOportionality exists between the heat of swelling of a binary aqueous system and the correSponding volume change. Since a purely theoretical determination of the function H* is not possible without simplification of the work function A'H, the decision was made to obtain an estimate of H* using the 58 empirical data of Stamm (1938). It seems reasonable, judging from the nature of the terms present in equation (2.4.26), that u* will not vary appreciably from one biological material to another. Thus, while the quantity (AH/AH°) - the ratio of the heat of adsorption at any P/Po value to that at saturation - characterizes the specific character of the individual substance, the deter- ministic function u* classifies it as a class or sub-class. The work sheet used for the calculation of u* from the empirical data of Stamm (1938) is shown in Table 2.1. Column 1 shows the standard relative vapor pressure and column 2 gives the corres- ponding specific adsorbed mass read from the isotherm. The figures given in column 3 are the mono-layer capacities determined by the standard B-E-T procedure. Columns 4 and 5 show the numbers of adsorbed layers of water molecules. Column 6 is simply a reduced form of column 5, obtained by dividing each value of column 5 by the number of adsorbed molecular layers at saturation. Column 7 contains the isosteric heat values for spruce wood taken from Figure A.l of the appendix. Column 8 shows the average AH values in terms of molecular layers, obtained by careful averaging of the relevant values of column 7. Column 9 expresses the average iso- steric heat values as a ratio of the isosteric heat value at satura- tion. Column 10 is a tabulation of the density values obtained from Stamm (1938), and columns 11 and 12 are reduced forms of column 10. Column 13 is obtained by dividing terms in column 12 by the corresponding terms of column 9 in accordance with equation (2.4.24). 59 000.H 000.H 000.~ can 000.0 m 0 000.~ 00~.H «mm 00.0 : 00m. 00.~ ~00.H ~00.~ 000.H «0m nmw. o % m~0.~ cad.“ «0m 0m.m : mum. m0.0 q~0.~ c~0.~ 000.H «0m saw. m * «m0.H and.“ own 0m.¢ : mum. 00.0 mm0.H N¢0.H 000.H mmm Hum. a * 0m0.H ~m~.~ 00m 0m.m : mca. 00.0 moo.a ~00.H 0N0.fl 00m 0N0. m ¢ 050.H «RH.H mom 0n.~ : mma. 0n.0 000.~ m~.~ «00 0N.N : mag. 00.0 m¢0.H H0~.H mm0.H mac mmm. N e 000.H ¢0~.~ 0H0 00.H : m00. 0n.0 OHH.H m-.~ 000 00.H : 000. 0d.0 0-.~ ¢MN.H one m~.H : moo. 0m.0 00.0 mmH.H qm~.a 0¢~.H 00«.H 000 #00 qu. H % 00.H : Ono. 0N.0 00H.“ «NN.H 0am on. : mac. 0~.0 05d.a omm.H 055 cm. : 0mH0. no.0 00H.H 00m.H .. 00.0 0m0. o. o . mm a... a a o «.1 So So .0 mm mum mo ax; mamas x? x m3 ma NH HH 0H m 0 h 0 m a N H .oomm um 0003 moauam MOw Awmmflv EEmum we oumv HmoHuHan msu Boom omcHEhouom mm *3 cofiuocsm mLH .H.N mHQMH 60 10.. 93.5.}.1 I55 I» I'rwWhHUW‘ 000 000.0 : 000. 00.H 000.H 00.H 000.H mom 00.H q % mom 000.0 2 000. 00.0 000 HNH.0 : ONN. 00.0 mq0.H mmo.H 000.H 000 05.0 m t 000 00¢.N : MHN. 00.0 0H0 00H.N : 00H. 05.0 00H.H 000.H 000.H 000 00.0 N * 0N0 0N0.H : 00H. 00.0 #00 ~00.H : 00H. 00.0 000 05¢.H : 0NH. 00.0 005 HNN.H : 0HH. 00.0 005 050.H : 000. 0N.0 mmN.H 0m. NNm.H «0m 0N.0 H e NHO H05. : 000. 0H.0 ~00. E oo\a 43 .mm mm ma ”\ax mowaoa sx\x ax x om\m 1| x\x HH 0H m 0 . n 0 m o m N H .Aomoav moaoa soowaoooa Mo mumw Auoqv EumSuOmH COHuauOmow msu co vmmmm .cuoo pom - O . ... one 43 n Q\o ooaooaou ooooooo och .N.N massa 61 000.H 000.H 00.H 00m 000.H n 0 000 NNN.0 : 0N. 00.H ~00.H n00.H 00.H 00m n00. 0 0 0N0.H 0N0.H H00.H H00 N00 0H5. m 0 mom.0 : 00H. 00.0 000.H mmo.H mHo.H 0mm Hum. 0 0 000 500.0 : NMH. 00.0 N00.H 000.H 000.H 5H0 0N0. m 0 0N0 nn.N : 000. 00.0 000 0N.N : 000. 0n.0 HNH.H n00.H Nno.H 000 m0N. N 0 0N0 . 000.H : 000. 00.0 000 0N0.H : 000. 00.0 N00 Nnm.H : 000. 00.0 000 mH.H : H00. 00.0 00H.H 00. 00H.H 005 NOH. H 0 000 000. : N00. 0N.0 0H5 000. : 0N0. 0H.0 on“ 00. : 5H0. 00.0 5000. a 030 «1 oWW mm ma WWW” muwmma ax\x Ex x om\m HH 0H 0 0 n 0 m 0 m N H HoooHV EuosuomH coHuaMOmwm ago so wmmmm .oooooo toe x.:<\: R1). 1 2 The total volume adsorptively filled during this step is: R2 - v12 =.- J‘ vr dR [25.131 R1 74 It is the implication of the capillary condensation theory that: v = UR L(R) [2.5.14] 12 where the capillary is assumed to be cylindrical with mean radius ‘R = % (R1+ R2), and length, L(R). Making the simplified assumption that for each finite adsorp- tion step the associated hydro-expansion Av is restricted within the corresponding volume v it is justifiable to consider for 12’ suitably small steps that: vS = v12 [2.5.15] Thus, equation (2.5.12) can now be rewritten as: Av = 3(1-2v) v12 80 (Ma- Mo) [2.5.16] This volumetric expansion means, in effect, that a pore of mean radius R changes its volumetric capacity from v in adsorption to 12 v12 + Av in desorption. The volumetric hydro-expansion, as defined above, depends on the mechanical prOperties of the product. Since these prOperties exhibit water-induced changes of physico-chemical origins, their collective contribution to permanent hysteresis is well represented in Av. For purposes of usage, Av is converted to its gravimetric equivalent by introducing the apprOpriate density term, D, so that: _ AX = _ AMd - p 3(1 2v) v12 Bo (Ma- Mo)/p [2.5.17] 75 If the bulk modulus, K of a substance is known as a function Of moisture content, it can be readily shown that equation (2.5.17) can be written in the form: A M R T 1n(Po/P) gig . 2. . d va M K p [ 5 17a] 2.5 c. Double Superposition £5 "Capillary Condensatiog: - and - "Swelligg Fatigue" - Hysteresis USing the Boltzman superposition principle that the effect of the sum of causes equals the sum of effects of each of the causes, total hysteresis may now be approximated by the following additive procedure: (1) Compute for each Pa/PO value, the specific desorbed mass, Md, using the relationship: M = M + A M a 0 [2.5.18] d (ii) Compute, using the Cohan relationship, the desorption relative vapor pressure, Pd/Po, corresponding to each Md value, such that: w Pd/PO = (Pa/Po) where W is empirically determined for each product. (iii) Plot Md obtained in accordance with equation (2.5.18) against the corresponding Pd/Po value to Obtain the desorption isotherm. This scheme will be tested in chapter four for biological products whose bulk moduli have been empirically determined as a function of moisture content. 76 III EXPERIMENTAL The adsorption and desorption equilibrium moisture isotherms of pre-cooked freeze-dehydrated beef were determined gravimetrically by exposing the product to a temperature-controlled free water sur- face under vacuum. The sorptive weight change was automatically and continuously recorded by s Cahn electrobalance. 3.1 Product Preparation Low fat commercial grade beef was secured in the form Of approximately one-inch cubes. The initial fat content was approxi- mately 9%. The product was cooked in a forced-convection-air oven at 300°F for a period of 30 minutes. The product cubes were placed in aluminum trays and covered during the cooking process. The cooked beef cubes were subsequently frozen in aluminum foil trays at 20°F prior to freeze-drying. The freeze-drying was done in a commercial freeze-drier at a platen temperature of 105°F with a pressure of less than 1 mm of mercury. Both the platen temperature and pressure were held constant during the freeze-drying Operation. Approximately 15 to 20 hours were required to dry the product to less than 2% moisture content on the dry basis (d.b.). After drying, the product was ground in a Fritz-Patrick mill using a 0.063-inch screen. The resulting powder was immediately placed in sealed bottles which were stored in a desiccator at -20°F. Portions of the product to be used for equilibrium moisture content determinations were equilibrated to the isotherm temperature prior to the run. 77 The moisture content of the freeze-dried beef following the freeze-drying Operation and prior to preserve-storage was determined in a heated-air oven in accordance with a procedure prescribed by Triebold and Aurand (1963). In this method an oven temperature of 80°C was used for a time duration of 16 hours. 3.2 Determination of Equilibrium Moisture Content 3.2 a. The Adsogption System: Equilibrium moisture content of the pre-cooked freeze-dried beef powder was determined by observing the weight change of the product sample when exposed to a constant water vapor pressure under vacuum. Following the attainment of equilibrium at a specified condition, the vapor pressure was adjusted to a different value and the new equilibrium point determined. In this way it was possible to obtain adsorption and desorption isotherms in one continuous experimental process on the same product sample. The basic component of the system was a Cahn RG automatic electrobalance. This highly sensitive instrument (effective sensi- tivity of 0.2% of full scale) was enclosed in a Cahn glass vacuum bottle with hangdown glass tubes for product exposure; see Figures 3.1 and 3.3. The principle of Operation of the balance is illus- trated schematically in Figure 3.2. The instrument in essence, converts the sorptive weight change of the sample into an electrical signal by the use of a beam position sensing apparatus shown in the diagram. This signal is amplified in a 2-stage servo amplifier and then employed to drive a strip chart recorder. The electrobalance 78 < 1.. .3000 2:880 2 want: 3. u. .20 308.005 93>: u .4. 0. N. :m 9.250% 29.50.89 ....m mane... mum—24:0 200065 0041.0 2. moz<4mwmm 711 02:... 23:13 sass // . momzmm $2.56..sz 33.25023. , 30.2% 0.9.03. mmhi than )1 5.23 D mmomoomc m 3 11.3% [111. _ o “.5: zones _ .. an... [1, \1» ._.So a [\ 02: 2.52.. z.<...\~ .6802» 2.0: L 79 .mucmfimnouuowfimuczwo msu mo wHaHocHum .o.m mpsmww cuczoouc 292.933 .83... . fl .. _ . _ _ 0 8°;- ¢UFJ—E _ _ l— . 1 h 32:. 52.8%. 2. 32:. fl .9... 4......“ 4 L - 2. i.e.? 80 81 Figure 3.38 Close up view of the Cahn automatic electrobalance in the vacuum line. Figure 3.3b Control unit and recorder outside the experimental chamber. works on the null-balance principle. Thus, an accurate and con- tinuous recording of product weight change can be obtained in the entire sorptive range. The control unit (see Figure 3.1) functions as a standardization and calibration device. Figure 3.1 is a schematic diagram of the vacuum line. The glass chamber containing the‘balance was connected at one end to a 13 mm inside diameter main vacuum line, and at the other end to a pure water vapor source. The main line was connected through a vapor trap to a mechanical-diffusion pump combination capable of pumping the system down to a pressure of l x 10'3 torr in fifteen minutes. The vapor trap was a glass trap with external reservoir. In its position at the inlet to the pump, the trap served two pur- poses: first, it acted as a baffle to the molecules of the pumping fluid back-streaming from the diffusion pump, and secondly, it pre- vented the possible contamination of the pumping fluid by condensable vapors emanating from within the vacuum line itself. The vapor source consisted of a temperature-controlled dis- tilled water reservoir; F of Figure 3.1. The temperature of the reservoir was controlled to within 0.05°C using a constant tempera- ture laboratory bath with an effective Operating range of -30°C to +71°C. Thus, desired relative pressure values were easily attained by precise temperature control of the free water surface in the reservoir. As a check on steam table predictions, pressure measure- ments were made in the system by means of three over-lapping and independent vacuum guages, namely: (i) a mercury u-tube manometer 82 for measurements in the pressure range from S to 760 torr, (ii) a Virtis McLeod guage for the range 0.5 to 10 torr, and (iii) a standard type McLeod guage for the pressure range from 1 x 10"3 to 0.5 torr. All connections in the vacuum line incorporated hyvac step- cocks in such a manner that the isolation of any desired component of the system was easily achieved. The entire apparatus, with the exception of the control unit and the recorder, was enclosed in a 5 ft. x 4 ft. x 9 ft. controlled temperature cubical. Figure 3.3(a) is a close-up view of the balance in the vacuum line, and Figure 3.3(b) shows the control unit and recorder outside the experimental chamber. A continuous record of temperature was monitored by capper- constantan thermocouples from (a) the immediate vacuum around the test sample, (b) several locations in the cubical, and (c) the vapor- regulating bath. §;£__2, Procedure For the adsorption test, about 100 milligrams of the sample weneplaced in the aluminum sample container, and the entire system evacuated to l x 10"3 torr in order to establish the 0% relative humidity reference with which the sample was allowed to attain initial equilibrium. After the sample had attained the isotherm (cubical) temperature, the stapcock S4 of Figure 3.1 was Opened to expose the sample to a pre-conditioned free water surface in the reservoir, F. The adsorption isotherm was established by adjusting the vapor pressure of the free water surface to pregressively higher 83 values corresponding to higher relative humidities and allowing time for equilibration at each level. »A minimum of ten experimental moisture equilibrium points were employed to establish an isotherm. Following the establishment of equilibrium at 100% relative humidity the vapor pressure was progressively reduced - step by step - to determine the desorption isotherm of the same sample. 84 IV RESULTS AND DISCUSSION 4.1. The Determination of Pore Structure from Water Sorption Isotherms - Verification of a Pore-Size Distribution Function of the Power Law Type for Bio-materials The crux of the isotherm equation deve10ped in chapter II, is the power law type distribution of pore radii which was postu- lated. It is therefore of central importance to the considerations to be undertaken in the present chapter, that the basis of this hypothesis be explored and verified in detail. The determination of pore-size distributions of catalysts from sorption isotherms is a well-established technique (Lester, 1967). In this technique (See for example Gregg and Sing, 1967, pp. 160-172), the Kelvin equation is applied directly to the desorp- tion branch of the isotherm; the results so obtained give a qualita- tive picture of pore structure. Of the various refinements which have been prOposed (Carman, 1951; Barrett, Joyner, and Halender, 1951; Harvey, 1943; Dollimore and Real, 1964; Viswanathan and Sastri, 1967), the approach of Cranston and and Inkley (1957) is of particular interest because: (a) it is more exact than any of the other methods, (b) it may be applied either to the adsorption or desorption branch of the isotherm and (c) the total Specific sur- face area estimated by this method can be compared with the corres- ponding B-E-T values and the differences used as an evaluation of the physical assumptions regarding the pore geometry. The original scheme of Cranston and Inkley is generalized to handle pores of shape other than a cylindrical geometry as follows: 85 Consider a finite adsorption step from relative vapor pressure Pi/Po to Pi+l/Po with condensation occuring in pores with radius varying between R and R (P P R > R1). If the step in 1 1+1 1 ’ 1+1 relative vapor pressure is kept small, the above described pores -—- 31+ R1+1 have an average radius of R1 8 -——§-—-— . At the same time, smaller pores are already filled, while in larger pores the thickness > 1+1 of adsorbed layers on their walls increases from ti to ti+l° The total volume of water adsorbed in the step, i = l is shown in appendix B to be: R2 vl2 = IRI vR dR 71. IR2 (R-tl) = . f . v dR N. R 1 R °° R-t v 12 . f . a + (cz-cl) IRZ a 2 R as [4.1.1] where t12 = (t1 + t2)/2 VR = total volume of pores in the range dR K, fl and f2 are shape factors such that: {a a 2, fl n a 3, f1 = 3/4, f2 - 3 for a spheroidal geometry R-t and the quantity C-i-l) is a correction factor for the curvature = 1, f2 = 2 for a cylindrical geometry of the meniscus. Assuming that V is suitably constant over the range R1 to R R2 and letting this value of VR be specified as V 2, equation 1 (4.1.1) becomes: 86 + (t - t ) . f . dR 2 l 2 R2 2R2 Rearranging equation (4.1.2) gives: m R-t 12 v =r (v -k F . v .dR) 12 12 12 12 “R2 2R2 R R2 K n where r12 = (Rz-R1)/f1 . [R1 [(R-tl) /R ] dR U l/f1 [(R-t1)/R]K :51 —. for a suitably small difference, dR = (RZ-Rl) and R12 = 2f2(t2-t1) [4.1.2] [4.1.3] [4.1.4] [4.1.5] For computational purposes, the integral term of equation (4.1.3) is replaced by a summation term of all increaments of radii from R2 to the largest possible pore radius, so that: Rmax R-t 12 = r { v - k L ( . 12 12 12 12 R2 2R2 R V Pore-size distribution is defined in terms of: v . AR) ] [4.1.6] (a) Pore-Length L(R) in the case of a cylindrical geometry where: -2 L(R) = VIZ/W R and cpz(R) = m dR 87 [4.1.7] [4.1.8] Also, because V H R2L(R) 12 and $12 = 2 H R L(R) therefore 812 = 2 VIZ/R [4.1.9] or (b) Pore—Number N(R) for spheroidal "ink bottles" where: 3 4 .. N(R) = val-5 n 3 [4.1.10] 42m) = anon/an - [4.1.113 and 512 = 3 VIZ/R [4.1.12] Cummulative specific surface areas computed using equations (4.1.9) and (4.1.12) can be compared with corresponding B-E-T surfaces - thus, affording a convenient criterion of geometrical selectivity. The complete calculation is performed on a CDC 3600 computer using a modified Cranston and Inkley (1957) method. In this scheme only 12 to 15 experimental isotherm.points are read in as shown in the illustrative input tabulation of Table 4.1.1. The 40 or more points usually needed are synthesized by the use of the Aitken- Neville modified Lagrange iterated interpolation scheme (this scheme is discussed by Moursund and Duris, 1967, p. 135). After this point, the calculation of pore-size distributions follows closely the method of Cranston and Inkley (1957). The essential differences being that the "cylindrical" formulations of Cranston and Inkley are replaced by their generalized counterparts as developed earlier in this section. This generalization makes it possible to assume any desired geometri- cal model for a pore. Two geometrical models - cylindrical pores and spheroidal ink-bottles - are investigated for comparative reasons. 88 Table 4.1.1. Illustrative input data for beef powder freeze-dried at 105°F platen temperature. Isotherm.temperature - 10°C P/Po M; 8 mass va 0 adsorbed Pore Radius (R) 2 for PIPo adsorbed (gm/gm) (cc/gm) (Ac) 0.000 0.000 0.100 .070 0.0515 6.742 0.200 .095 0.0699 9.635 0.300 .110 0.0825 12.463 0.400 .122 0.0927 15.617 0.500 .135 0.1026 19.680 0.600 .155 0.1178 25.624 0.700 .191 0.1546 34.563 0.800 .260 0.2234 52.743 0.900 .380 0.3375 106.054 0.950 .450 0.4240 214.982 .963 300.000 1.000 .505 0.505 89 It turns out that in all cases investigated (see Tables 4.1.2(a) and 4.1.2(b)), the cummulative specific surface areas calculated on the model of interconnected spheroidal ink-bottle geometry com- pares better with the B-E-T surfaces, than the surfaces bases on the cylindrical pore model. It is concluded on the basis of this finding, that the pores of biological materials are best modeled for theoretical treatments as interconnected spheroidal ink-bottles. Figures 4.1 through 4.4 are log-log plots of pore-size dis- tributions of several biological materials calculated on the basis of the spheroidal ink-bottle pore model. The remarkable linearity of these plots indicates that a strong tendency exists in bio- materials toward a power law type distribution function for pore- size. Thus, the power law equation (2.3.16) when expressed in its logarithmic form: In m2 = In K + y In (R) [4.1.13] 1 yields an empirical straight line with a SIOpe y, and an intercept of magnitude 1n K1. In Table 4.1.3, the parameters Y and K1, as determined from the pore-size distribution plots, are summarized. These values were obtained from least square straight line approximations of the data plotted in figures 4.1 through 4.4. In spite of the elaborate refinements of the Cranston and Inkley method, there exists (as with the other methods) no generally accepted accuracy criteria for evaluating pore-size distributions so 90 cwo.o ammo. anon~.m o~o.o smoo. moo. no. ass. sss. ssm.s .4.s. .-o.xs.. s.s.s osoo. boo. sa. sss. sss. 4e4.4 sass. .-o.xo.~ sss.4 secs. soc. ms. 4mm. ~4s. s4s.s ss.s. ~-o.xs.s 444.4 osoo. .oo. ms. .ss. s4s. sss.~ so... .-o.xs.. sa..s woos. soc. mos . sss. s4s. .4~.~ .sos. .-o.xo.m 4ss.. «soc. Nos. 4.. sss. ass. 4so.~ was.. N5...... oo4.. swoo. Nos. s4. .o4. sss. ess.. sass. .-o.x4.. ssn.~ s4oo. 4oo. sss s04. 4ss. 4.4.. .Ns.. ~-o.xs.s 4s... smog. soc. s4. so4. ens. oss.. some. .-o.xs.. ass.o ..oo. .00. ms. sO4. ass. 4o~.. sons. ~-o.x~.s smo.~ s4oo. 400. so. «.4. sss. snow. 4444. ~-o.xs.s 4.4.. 4soo. moo. mas s~4. oes. «.44. 4nss. N.s.xs.s ss..~ ssoo. moo. sss os4. «as. mass. sass. ooo.o os~.. 4soo. soc. nss ss4. nos. am”? .s knee 3|... . £428 . flaws“... .58. .Maauwmsn fin... . . . . u . . . 4. . . 4 .4 o us go a CO ‘Uflfia “ 1H” NC“ wwawn =0 ”00“ H+H>Od>l HON UUQHOGQU ”PDQ AVUUHHOQNUUGWV ~.m w ~.s .4494 .44. s.> ~.> .as.o> amunuse .4. s s s 4 4 s 4 s N . .vuasaus a. sausaoom anon Hooaumcudmo a .auonuon. 0.0. so mouse was .oaunon hoaxcu one coueasuo moauqmoa a an mousnsoaso .ousuousaaou saunas memo. an mmqumamnmouu nausea mean you acuusnauuauv oaaocuuom ~m.~.~.4 magma 91 Odd. mmo. ¢Q.NO HO.MH OHKO.Q on.hmc “duo. 5H0. m.Nm hnH. nmo. MO.QQ ma.- OHXN.H 0O.m¢m néNo. OHO. n.5n cud. mun. Om.dh n0~.w OHXm.¢ ON.ONM nwuo. mac. n.Nd Had. nnb. h~.no ON.O~ on~.n OO.H¢M memo. ado. n.h# OaN. ssh. 50.Nn ohh.nu nhm.h Oh.wmn osmo. Ono. mm OfiN. has. o~.an ~o¢.h OHKN.N OQ.NO~ M¢No. ONO. no new. ONO. #5.Hm 00¢.o Nom.¢ ON.NM~ NdNO. ONO. mh MON. Onm. n2 ON.mN own.¢ m¢h.¢ ON0.0w MONO. mac. mm 0, 00m. wow. m¢.ON ONN.¢ OON.~ OHN.HN nONO. sac. mm ham. Hmw. CNN.@~ NbO.H mn~.¢ omm.mN ¢O~o. OOO. meg cum. mam. O¢N.¢~ mwo.m MBN.H ONN.N¢ sumo. mHO. med H¢m. NCO. On~.~H nNOm. mm~.m Ohm.H~ onoo. moo. mmfi o¢n. OHO. oo~.o~ O¢NO.N o~n.~ 05¢.QN OfiuO. NHO. nma wan. Ona. mmd.m ONHO.. «NN.H mNh.o #QOO. «O0. m¢~ Non. NNO. nun.h Ame.” H10~x0.w ONm.n~ ONdO. OHO. mmH Nun. 5N0. m w m o m d m N n O A—u. UGOUV ...~...4 ..... .oaasom mucouomu.m “causes a so >-muwuwasc econ no: co.um.ucououm.m « 4n... n mus a Inseam m hm. «Haw Hum AoooH um >u40somo mommanocoa co vmmmnv Awavmmm I m moo. o.4~. mmo. o.4 H4o. mmo. o4mo. m.o. n.m 4mo. mo.. o.mo. moo. m.o moo. .ma. mmoo. 30. mg % moo. smH. smoo. moo. o.m mso. ~4N. n4oo. ooo. o... .oo. Hun. omoo. moo. o.m. moo. omm. o.sm. woo.s No.XM.s oo.44s omno. ooo. o.m~ moo. 444. m.om~ o4o.o o.x..o oo.H~o 4omo. ooo. 0.5. moo. ~m4. «.44. o.~.m o.xm.m o..mmo omso. soo. o.o. «0.. «no. o.ona o4.m. o.xo.~ o.osm. mHNo. wHo. n.- s on.. ass. m.o~. mo.m. ono.n o.~oo. smuo. o~o. n.n~ o o s o m 4 m N H u Av.uaoov AmvN.~.4 manna 94 Ohm. 0mm. mm¢.0 5mm.a OHIOHKm.H OHIOHKO.M 0000. 000. mwd 40m. N40. OOO.N 0mm.a NAIOHKO.O ONIOHx0.n Mano. BOO. man ~0m. m4o. MOH.0 ~H4.H HH:O~KO.H OHIOMKN.N 0000. 000. MON hmn. N40. Anh.4 OON4. nH10H#M.O HNIONK4.N HMOO. N00. nHN 00m. OmO. ««n.4 «o.4. «.-o.x«.n ..-o.xs.4 .soo. «as. s«« #04. Nnm. «.s.s 4sos. «.-o.x4.4 o.-e.x«.. secs. 4so. ss« m04. 4n0. «o..s sass. «.-o.xs.« ..-o.x..n .404. «so. s4« N04. 0mm. ««s.« s4«.. «.-o.ss.m ..-o.x«.« s.oo. .oo. ms« 004. “no. s4s.« «s.«. «.-o.xs.« ..-s.x..s secs. 4oo. s4« NNQ. mmo. 0mm.“ mnfim. NH10MNN.N amuoux¢.m 5400. MOO. th mNQ. 000. num.~ NNMO. NAIOHKO.M HHIO~XN.0 0500. nOO. MON Om4. N00. HN04. ané. 000.0 4|O~x#.4 N400. MOO. mmN mm4. mom. A BC A5\UOV .«a 5...... as... .348. 3.4.. 00356 .20.... do amt—.40 3 £40.20... “+0 4> newswoousuaa .3350 I so nouns melm u 20» so woman 1 cu A440 so mouse moAHOuvs na.mmu «.8 «.m 9.0«9 2:. «.> «.> 93.2. 32. .4322 as}. O O 5* 0 n 4 n N a .4234: a. House 32. 0.33-4.3 .3885? 4 .auonuon« 0.0. on mouse was .oauzun hasacu mas couucsuo magnumoa a he nodumusosso .ousuauooaeu saunas mono. us veaumuouomuu somaoo «use msxooououo new mouusnuuunuv unannoum NAON.H.4 odoufi 4.42 84 .4 4.2.5. «-4.x«.4 .«8. 84. 4 . «« 44.. 444. o..4« 4«4.« 4-4.x..4 «-o.x4.. s4«4. «.4. 4.«4 «4.. 444. «.4«. 4«4.« 4-4.xs.« «-4.xm.. 4«44. s.4. 4.«4 4«.. 4««. 4.44. 4.«.. 4-4.44.« 4-4.44.« s4«4. 4.4. 4.«4 .s.. 44«. 4.44. .4..« 4-4.4s.. 4-4.44.« «444. s.o. 4.«4 o.«. «««. «.... 44s.« 4-4.x«.4 4.4.4«.« 4444. 444. 44 o4«. «s«. .4.«4 444.. s-o.44.4 4-4.xs.« 4444. 4«4. s4 44«. 4«4. 44.44 «44.. 4.4.44.. 4-4.4s.. .444. 4«4. 4« 44«. 444. 4«.44 444.. 474...... 4.4.4... 44«o. «.4. 44 «H 444. 444. 4..44 so..s o.-o.xs.. s-4.xo.4 44«4. «.4. 4s «.4. .44. 44.44 .44.4 s-o.x4.4 s-o.xs.« «4.4. s44. 44. 4«4. 4s4. 44.44 s44.4 o..o.44.s s-o.xs.4 44«4. 4.4. 4.. .44. «cs. .4.4« 44«.. o.-4.x«.4 o..4.x..s 4«44. 444. s«. 444. 4.s. «4..« 4.4.4 4.-o.x..4 s-o.xs.. ss.4. «.4. 44. 444. 4.s. .«.«. 44«.. o.-o.x..4 o.-4.x«.4 4444. 444. 44. «44. ««s. 4s.4. 44..4 o74.44.. s-o.u... 44.4. 4.4. 44. ««4. ««s. .4.«. ««4.. o.-4.x..4 o.-o.x«.4 4«44. 444. 44. ««s. 44s. s4... 44s.. o.-o.xo.. o.-o.xo.4 4..4. «so. 4«. s 4 « 4 4 4 4 « . - .4.44444 .44«...4 4.444 .mao£04 mucoumum.u .muuaoo 4 4n 4..4u.uua:a mean 443 ao.uq.uamuomu.o « Hum . n «mm a m 44 4 4.4«4 44. no.0. an 4u.uuauo 404$.uocoa co vomunv 4464 «an a swam woo. 440. 0.4 .40. 440. 4.0. 4.4 440. 40.. a8. m.o 6 440. .4.. o, 400. 4.4 440. 44.. 400. 0.4 440. ~44. 400. 0... .40. .44. ~00. 0.4. 440. can. 400. 0.4. 440. 444. 4.444 044.. 4-0.x4.. 4-0.x4.¢ 4000. 400. 0.4. 440. 444. 4.44m .00.m 4-0.x0.. 4-0.x0.~ 4440. 400. 0.4. 40.. ~44. 4.44m .40.~ 4-0.«4.w 4-0.34.4 ammo. 4.0. 4.44 04.. 404. m m 4 4 4 4 4 N . - .4.44444 .444...4 4.444 1n [$2(R) = dN(R)/dRJ 97 '7 ' :L: o «r- dl- o q- o 4.. J- 43 D O ‘ .A o ' f o o o A t o ' A 4. 44 i A0 4; ° .Jf‘ t ) 0 6 -—--+- -+-»+~H+++n— b. n D»o a n T -:L (3 A6 Figure 4.1 Pore-size distribution plot for corn at the tempera- tures: 4°C, 15.5°C, 30°C. Q3 Based on the desorption data of Rodriguez-Arias (1956). w~+-+~+~+ H-H+ m 0 'érl o oo 4‘ .A on .13 A ° T 9°“ + I - i O ++~~-+—~—+~-——-~-++++++—+—-+ ..+ - mm P 4 4 c 4 £5. :02 IO l.0 nnnt: nAn-rivn 1.- in [A 0‘) ' i1.‘.j.fll|.fv: :~ 1 :¢2(R) = dN(R)/dRJ In 98 6. 54. —+ W'WWW‘HW D O U o D 0 a pm 0 ‘9 ON» a p o 0 l8 l2; 0° .3 n o 30 C A A-—- A no ° ‘ 0 H3 0 ° Figure 4.2. Pore-size distribution ‘P plot for cotton at isotherm 9 A. temperatures: 10°C, 20°C, a Q 30°C. Based on the adsorption A o a data of Urquart and Williams (1924). ‘II A o l0 '1: 5,4“ I: ”.A ‘D a -- 8'0 #- Ago. 0 o .4 O n 3 4 ’IZ. .. ----.. . - - .. +4-} 1 .4 no 33+H+1 r ffiHH—r + 1344 s ' PORE RADIUS — 1n iR(A°)J I67 -m..-+n_+..+—+_T_f-m H3 -—+++1+'r++----+-—'r—~t—+~++—t+r— 52. ~ +~+~+4+++H9——- + + ~+—+—++—++H—--+~ 5. Ié'gl-H—Hw 1 I03 99 u o a a o A u n n o o A A “A o 8 a 20°C A. ° . °-_- 0 a 60°C a o n— n u 0 100°C 0 _ A A 8° A A .0A Figure 4.3. Pore-size distribution 0A plot for wood at isotherm 0 temperatures: 20°C, 60°C, 100°C. 0 Based on the adsorption data of an OA Lewis (1921). 9e. A J’ nI: 0A A D o a A %o 0 4% 8 0° 0 A +- i—vwggHH-aw w» +H+l I s . i 4 l l() L() PORE RADIUS - 1n LR(A°)J 742m) = dN(R)/dR] b— 1n IO 4. -r ‘llllll - +~++~H+++--~ +- -~+ "a. + -+-++++H— ~ -I .9- I6" --+- ~+—+-H++H- ~—+ -+- -r ++++++ a... 5.; N 6 0 l + 100 go A A o o D O o£4 a” o o A 10°C o __ A o n A 22.2°C a _ ° 0 .A o 37. C Av——-°7o G .A o n 0 A Figure 4.4. Pore-size distribution a n A plot for pre-cooked beef powder 0 freeze-dried at 105°F platen a An temperature. 10°C isotherm, oo n on 22.2°C isotherm, 37.7°C isotherm. °o 0 Based on adsorption isotherms 0 AA obtained in the present study. A 6' o o 48. a A A '5; .A o A a a? m9 D ~I+H+H+~+ -—--+ + LO T .L fiW—f |PORE RADIUS - 1n IR(A° I)J Table 4.1.3 Summary of the characteristic parameters of pore structure as determined from pore-size distribution plots. PRODUCT Temp Y K1 0 § from from = 4 + Y = 4/3 0K1 °C pore-size pore-size EQN. EQN distribution distribution ' plots plots (2.3.20) (2.3.19) 4.0 -4.636 .335 -0.636 1.404 CORN 15.5 -5.015 .237 -1.015 5.182 30.0 -5.059 .853 -1.059 7.763 10.0 -4.394 .049 o0.394 0.207 COTTON 20.0 -4.560 .214 -0.560 0.897 30.0 -4.603 .238 -0.603 0.995 20.0 ~4.416 .0934 -0.416 0.391 WOOD 60.0 ~4.555 .2053 -0.555 0.860 100.00 -5.202 .448 -1.202 18.634 PRE-COOKED 10.0 -4.344 .119 -0.344 0.497 FREEZE-DRIED BEEF 22.2 ~4.599 .487 -0.599 2.039 POWDER 37.7 -4.876 0.756 -0.876 3.167 101 calculated (Everett, 1958). It seems apprOpriate that until such criteria are established so that the magnitude of the various quantities involved can be ascertained in absolute terms, pore-size distribution plots of the type represented in figures 4.1 through 4.4 should be considered as having qualitative but not absolute significance. Thus, the pore-size distribution parameters - Y and K1 - of Table 4.1.3, cannot be expected to be quantitatively accurate. Nevertheless, the exponential nature of the pore-size distribution function,‘Pz(R), is conclusively established from a qualitative stand- point, because of the well defined linear tendency of the log-log plots. In addition, the following observations regarding the para- meters y and K are qualitatively discernible from Table 4.1.3. : l 1. As characteristic parameters for the pore structure of the sorbent, Y and K are unique for each sorbent 1 system. 2. Both Y and K are clearly temperature sensitive. This 1 implies that the pore structure of the sorbent changes with isotherm temperature. While it is as yet premature to determine from a priori considerations, the exact nature of this change, it appears reasonable, in view of the effects of temperature on swelling forces, that the pores could undergo a certain amount of deformation with changing isotherm temperature. Although a purely tem- perature-induced expansion or contraction cannot reasonably be expected to alter the pore structure appreciably, 102 Hammerle (1968) did show that the so-callcd hydro- expansion in bio—materials is considerably temperature sensitive. In closure, it has been shown in this section that (a) the pores of biological materials are best modeled as inter-connected Spheroidal ink-bottles; (b) the pore-size distributions of these products are well described by an analytical function of the power law type; (c) while the present state of the art does not allow for precise quantification of the pore-size distribution parameters, useful insights have been gained into their qualitative characters. These results have important application in the theoretical considera- tion of moisture transport in biological materials. 4.2 Experimental Results Adsorption and desorption data for pre-cooked freeze-dried beef powder obtained in this investigation are summarized in Tables 4.2.1 and 4.2.2, respectively. The specific adsorbed mass is recorded on the dry basis (d.b.). Each value represents an average of three independent tests. The isotherm data are reproduced in Figures 4.5 and 4.6. The plots show the traditional sigmoid shape expected of porous water-binding biological materials. They are temperature dependent, showing decreasing adsorption with rising isotherm tem- perature. A distinctively pronounced hysteretic behavior is apparent in figures 4.7(a) and (b). While the desorption curves show a 103 < A D S O R P T I O N ;> in (d.b.) P/Po ‘1919 22.2°C 37.7°C .05 .047 .03 .0136 .10 .070 .045 .0209 .20 .095 .062 .0309 .30 .110 .075 .0391 .40 .122 .087 .0500 .50 .135 .100 .0591 .60 .155 .125 .0791 .70 .191 .163 .102 .80 .260 .230 .141 .90 .380 .33 .200 .95 .450 .405 .241 -98 .049 .465 .27 1-00 .505 .490 .286 Table 4.2.1. Equilibrium moisture content of pre-cooked freeze-dried ground beef. platen temperature = 105°F 100-150 milligrams 9.8% sample size Fat content 104 105 .50- .45. Io’c 40* 22 2 C 2 .35. 6 '2 .30. E 2 377.6 8 .25L UJ (I :3 5', 20+ 5 2 .154 .10; .05» Figure 4.5. Adsorption isotherms for pre-cooked hoof powdvr freeze-dried at 105°F platen temporaturv. (10”C - 37.7““) < D E S O R P T I O N -“—> in (d.b.) P/Po '19:g 22.2°C 37.7°C 0.05 .065 .04 .0427 0.10 .088 .057 .0500 0.20 .110 .075 .0546 0.30 .123 .085 .0682 0.40 .137 .097 .0846 0.50 .156 .113 .109 0.60 .191 .143 .138 0.70 .260 .200 .175 0.80 .425 .280 .213 0.90 .487 .465 .254 0.95 .498 .4755 .273 1.00 .505 .490 .286 Table 4.2.2. Equilibrium.moisture content of pre-cooked freeze-dried ground beef. platen temperature = 105°F sample size = 100-150 milligrams fat content = 9.8% 106 107 .s..? .451- .40- a 15 .351- 10°C ’5 u: 30% 22.2°C *z' o 37.7 c t) 25+ 111 a: .2 6 .20” 2 .I5» .I04 .05 0.0 .l .2 .3 .4 .5 .6 .7 .8 .9 1.0 PIP. Figure 4.6. Desorption isotherms for pro-cooked hoof powder freeze- dried at 105°F platen temperature. (10°C - 37.7°C) 108 o——— Adsorption .5()r °-—- Desorption (db) CONTENT MOISTURE 00.1.2.3; 5.8.7.8910 PH; Figure 4.7(a). Plot showing sorption hysteresis in pru-conkod freeze-dried beef powder. 109 .50~ u/ q’ .454 I, I I .40- I .2 I ‘-’ I . I 3 .35? desorpnon—jf i 1— i I 5 odsovplion 8 ‘ 1’ g / w .25;- / g I t- . / 'fl . / 2 i / . / / ' ISI- 15/ g / / . ; ,1 101' In/ ° I ’,n" W 2 .°" ? f// . 05;. J }/ II Figure 4.7(b). Plot showing sorption hysteresis in pre-cooked freeze- . dried beef powder (22.2°C isotherm) flattening tendency near saturation vapor pressure - suggesting a type IV classification of the Brunauer type - the adsorption curves are strictly sigmoid-type II, with the "B" point fairly well defined. These salient features sUpport - as do most sorption isotherms of biological materials - the micrOporous assumption so fundamental to the deve10pment here presented. While a number of biological materials will be used to test the theory as a means of demonstrating its SCOpe, the data presented in this section will in the main, be employed to test the generalized model. 110 4.3 Verification of the Derived Density Function for Sorbed Water It is assumed in the interest of simplicity that the B-E-T surface area (S ) of a given sorbent can be stretched out into B-E-T an equivalent flat surface. The mono-layer and subsequent multi- molecular layers are assumed to form a matrix, see Figure 4.8. The individual molecules file into vertical columns. The molecules of each column are forced by virtue of their presence in that column to conform to the cross-sectional area (oh, [A°]2) of the first (mono- layer) molecule of that column. The area, om, which one adsorbed molecule would occupy in the completed mono-layer depends among other things, on (a) the sorbate surface tension and (b) the intensity of the sorbent-sorbate interaction represented by the so-called spreading pressure at the surface. Because of the recognized com- pressive effect of this surface force together with its fundamental polarizability, the adsorbed water molecule is visualized in Figure 4.8 not as a spheroid, but as an ellipsoid, the degree of ”flattening” of which is prOportional to the gross compressive force acting on the molecule. According to the potential theory, this flattening is expected to be highest in the mono-layer and diminishes progres- sively outward in the succeeding layers. With the assumptions enunciated above, the multi-layer thick- ness t, can be defined as: t(A°) = va/sBET [4.3.1] 111 I T t _I_ Figure 4.8. To 1 I T2 TI <7'2"“(1'0 3‘1 specific surface strvtvhnd our Vertical section of stretched out specific surface, showing the multi-layer matrix. Each adsorbed water molecule is visualized as a flattened elipsoid reflecting Ihv degree of com— pression of the molecule as a result of intermolecular forces of attraction and polarizability effects. Each layer of molecules is subject to a different but uniform compression. The first layer being the most compressed and the tOp laver the least compressed. 112 The average thickness of a single layer of adsorbed molecules, T, previously defined by equation (2.3.7), becomes: va Ma T = t/n = = ————-—' [4-3-2] “ “SBET p “SBET where n is the number molecular layers adsorbed and p is the mean density corresponding to n. In accordance with equation (4.3.2), the average thickness at saturation vapor pressure, To is given by: M a T (A°) =-—————-— [4.3.3] 0 ponSBET Combining equations (4.3.2) and (4.3.3) yields: T E a: ;2. L4.3.4j 00 11 Equation (4.3.4) implies that the density ratio and the thickness ratio should be approximately equal. This furnishes an approximate verification of the density function derived, tabulated and graphed earlier in chapter II. In Tables 4.3.1 through 4.3.4, the above verification is performed for a number of biological products. A comparison of columns 5 and 9 of each table shows a remarkable agreement between the ratios. Three things need to be pointed out immediately regarding Tables 4.3.1 through 4.3.4. First, the saturation density, p0, is assumed to be 1 cc/gm, corresponding to the normal bulk liquid density of water. Secondly, since the B-E-T surface area, SBET’ determined from H20 isotherms is known to vary with temperature 113 Table 4.3.1. Isotherm temperature 10°C S Product = Pre-cooked freeze-dried beef powder. 2 Verification table for the density of sorbed water. 1 2 3 * 5 6 7 8 9 17/130 Ma _;5_ p/Do Va t Tn :2 m ' Mtge) ,, :61 a :5 = t/n Tn m = .091 p 5881‘ .05 .047 .52 .10 .070 .769 1.360 .0669 2.079 2.079 1.351 .20 .095 1.044 .30 .110 1.210 .40 .122 1.341 .50 .135 1.484 .60 .155 1.703 1.316 .1380 4.320 2.160 1.300 .70 .191 2.10 1.164 .2340 7.200 2.400 1.170 .80 .260 2.86 1.126 .3230 10.000 2.500 1.124 .90 .380 4.18 .95 .450 4.95 1.062 .4280 13.250 2.650 1.060 1.00 .505 5.55 1.000 .5050 16.860 2.810 1.000 * In these calculations, Po 114 1 cc/gm., the normal density of bulk water Table 4.3.2. Verification table for the density of sorbed water. Product Soda boiled cotton (Urquhart and Williams, 1924) Isotherm temperature = 10°C. SBET = 126(m2) 1 2 3 4 * 5 6 7 8 9 P”. Ma 3’} n 9’9. V. t Tn ;2 m n g .036 a ”(g-3°) va/SBET .05 .017 .48 .10 .023 .644 .20 .032 .896 1 1.220 .0295 2.300 2.300 1.217 .30 .041 1.15 .40 .049 1.372 .50 .058 1.624 .60 .068 1.904 2 1.120 .0642 5.000 2.500 1.120 .70 .080 2.240 .80 .099 2.970 3 1.089 .0990 7.800 2.600 1.080 .90 .132 3.697 4 1.035 .1390 10.800 2.700 1.037 .95 .164 4.593 5 1.008 .1780 13.715 2.743 1.02 1.00 .240 6.722 6 1.000 .2400 16.980 2.830 1.00 * In these calculations po ‘3 1 cc/gm., the normal 115 density of bulk water. Table 4.3.3. Verification table for the density of sorbed water. Product = Spruce wood (Stamm, 1938) Isotherm temperature = 20°C. SBET = 177(m2). 1 2 3 4 * 5 6 7 8 9 1”? “a 3’} n 9’9. a t n :9 m = “*(950) = M /0 _ Tn Km 2 .05 A“ a v /s - t/n a BET .05 .013 .26 .10 .025 .50 .20 .025 .99 1 1.15 .0434 2.45 2.45 .14 .30 .063 1.26 .40 .080 1.6 .50 .095 1.9 2 1.10 .0900 5.2 2.6 .08 .60 .113 2.26 .70 .135 2.7 3 1.067 .1400 7.98 2.66 .05 .80 .165 3.3 4 1.042 .1910 10.8 2.70 .04 .90 .225 4.5 5 1.035 .2410 .95 .275 5.5 6 1.000 .3000 16.8 2.8 .00 1.00 .340 6.8 7 1.000 .3400 19.6 2.8 .00 * In these calculations, 00 a 1 cc/gm., the normal density of bulk water. 116 Table 4.3.4. Verification table for the density of sorbed water. Product = Corn kernels (Rodriguez-Arias, 1956). Isotherm temperature = 4°C. SBET = 281(m2). l 2 3 4 * 5 6 7 8 9 "P. M. 3.;— n 9’92“ V. t T. _:_9_ m = “*(Zfi-O) . M [p -= t/n n Xm=.0865 a va/SBET .10 .065 .751 1.295 .0667 2.37 2.37 1.31 .20 .0925 1.07 .30 .1100 1.271 .40 .1275 1.473 .50 .1425 1.647 .60 .1625 1.878 1.145 .1510 5.370 2.685 1.150 .70 .1825 2.109 .80 .2125 2.456 1.043 .2480 9.00 3.00 1.02 .90 .2700 3.121 .95 .3000 3.468 1.000 .3460 12.4 3.1 1.00 1.00 .3500 4.046 * In these calculations, Po 117 211 cc/gm., the normal density of bulk water. (because the mono-layer capacity, vm, varies with isotherm tempera- ture), the absolute Specific surface area of the sorbent is not known. Even though it is customary to use a mean Specific surface based on several isotherms, this average value is questionable when wide variations in vm values occur with isotherm temperature. Because of the doubt in the SBET value used, the t (column 7) and Tn (column 8) values of these tables do not have absolute but do have relative significance. Consequently, the ratio TO/Tn (column 9), is unaffected by the degree of accuracy of S Finally, the veri- BET’ fication performed is insufficient as a test of the validity of the derived density function. At best, it only adds a certain margin of confidence to the derived function. The absolute test for the den- sity relation can be done experimentally by comparing gravimetric sorption data with the corresponding volumetric data for the sorbate, water, on the same sorbent system. There exists in current litera- ture a glaring lack of sufficient data along these lines and thus such a test is not possible at the present time. 118 4.4 Verification of the Derived Isotherm Equation 4.4 a. Theoretical Isotherms For illustrative purposes, two sets of isotherms calculated with the volumetric equation (2.3.22) are presented in Figure 4.9. The Specific adsorbed volume, Va’ is plotted in its reduced form by dividing each calculated isotherm point by the corresponding calcu- lated adsorbed volume at P = 0.98 Po' This reduction of all isotherm heights to l or 100% allows for a better comparison of characteristic forms of adsorption curves and facilitates the eventual matching of theoretical with experimental isotherms. The adsorbed volume at a relative humidity of 98% is chosen as a relatively well defined point. In contrast, considerable doubt still exists among investiga- tors (see for example, Bangham and Sever, 1925) as to whether equili- brium is ever truly obtained at saturation. As a result, the adsorbed volume at 100% relative humidity is not well defined. In these plots, the volumetric isotherm expression is chosen in preference to its gravimetric equivalent because it lends itself well for consideration on a general basis. The isotherm temperature used for the calcula- tion of Figure 4.9 is 10°C. Two sets of theoretical curves are pre- sented to illustrate the effect of the Pm value on the isotherm shape. The curves are obviously very characteristic showing (a) well defined initial "knees" at moderately positive values of n [the knees are clearly better defined at lower values of Pm, see Figures 4.9 a and 4.9 b], (b) certain linear parts, and (c) distinctive individual starting angles of the isotherms. All regular types 119 .0 E “O E momwm u oooa u e mAH.oV u a momma n oooa u e mamoo.v u a .ANN.m.NV .cem cuss .ANN.m.~V .cem suns “63350.30 mechanic: Hmoauowooafl. .Anvoé 35m: woumasoamo mEmnuOmH Hmoeuwuomne .Amvmé muawwm .68\a . o... m. m. N m. n. w. . N. e. o o . . _ a . . u . . 038 0.0209029“?an ":0. <2 a e “:“Q'Q s e 9: [95 "WM-707w ow mamas HUHJSIOW ason (D [‘5 '"d’djow/ 0w mamas 3301570717 03700038 120 of isotherms of the Brunauer classification are obtained merely by the variation of n. This finding is significant in that it vindicates the leading premise of the generalized theory that the three basic concepts in adsorption are complementary and thus can be unified into a self-contained isotherm equation. The present results must also be viewed in their broader context as strengthening considerably the significance of a purely power law distribution concept for the pore-size. 4.4 b. .3 as the Primary Characteristic Parameter of the Pore Structure As the primary shape criterion for the set of theoretical isotherms represented in Figure 4.9, n must be viewed as the character- istic value for the representation of the quantitative behavior of a porous sorbent at any relative vapor pressure. In the range n < -l, the theoretical isotherms are clearly of the Langmuir-type. Thus, this range of n values may be associated with materials characterized by exceedingly narrow pores, probably not more than two molecular diameters in width. These are the so-called micrOpores. Sigmoid type II isotherms are defined in the range -1 S 0 S + 1. Accordingly, pores so characterized fall within the so-called intermediate and lower macrOporous range (Brunauer, 1945; Gregg and Sing, 1967). n values in the range n > + 1 define type III isotherms and in conse- quence may be associated with large pores of the upper macrOporous range. It is well recognized that the isotherms of most biological products are sigmoid. Consequently, it would seem logical that the main focus of the present investigation should be restricted to the intermediate range, -1 S n S + l. 121 As defined by equation (2.3.20), n = 4 + Y [4.4.1] Ordinarily, this relationship would have provided a convenient com- putational formula for 0. However, it has been previously argued that the pore-size distribution parameter, Y, is not defined with sufficient quantitative accuracy to warrant this usage. This short- coming notwithstanding, the definition of n as a function of Y in equation (4.4.1) provides a preliminary insight into its qualitative character. It is unique and specific for each sorbent system. It is temperature sensitive, assuming progressively changing values as the isotherm temperature increases [see table 4.1.3]. As a conse- quence of the previously discussed relationship between 0 and the characteristic isotherm types, it is apparent that an appreciable temperature-induced change does take place in the sorbent pore struc- ture. Even though the exact nature of this change cannot be specified at this point, it is obvious that in applying the derived isotherm equation to specific products, 0 must not only be defined for each product, but needs also to be defined for each temperature. 4.4 c §_2_ a Secondary Characteristic Parameter g; the Pore Structure According to equation (2.3.19), . 4 7; == '5 TI Kl [4.4.2] g is thus a quantity combining the pore-size distribution parameter, K , with the factor of pore geometry, (% W). Since 5 is dependent 1 on K its value is unique for each sorbent system and it is also 1’ temperature sensitive (see Table 4.1.3). Because K1 is not accurately 122 defined from a quantitative standpoint, the exact value of 5 cannot be determined from equation (4.4.2). A progression from smaller to larger values of 5 is evident as the isotherm temperature is increased (Table 4.1.3). It would seem therefore, that a decreasing n value combined with an increasing § value implies a change in the pore structure. This change means that the effective size of each associated pore is affected by a combined hydro and thermo expansion or contraction of the solid component of the sorbing tissue. 4.4 d. Prediction of the Adsorption Isotherms of Certain Biological Materials with the Generalized Isotherm Equation Since it has not been possible to determine reliable values of n and g from a priori considerations, these quantities will be defined empirically to satisfy qualitative criteria stipulated from earlier theoretical considerations. To determine n for a Specific product and temperature, the empirical isotherm in its reduced volu- metric form is superposed on a set of theoretical isotherms such as shown in Figure 4.9. The best matching n and Pm values are thus selected. Based on these values of n and Pm, the function [Ejfi—fiflj is defined for each relative vapor pressure. Combining the approxi- mate density term, pl, with the above function, the best fitting average 5 value, E, is defined. It is now possible to define for each P/Po value, the first approximation of the specific adsorbed . ._ Zn.- AT] mass, defined by p1.§ [__0_—_ J. The discrepancy between this approxi— mate value and its empirical counterpart is minimized by the careful correction of the density term. This gives rise to a new set of density values, designated as oz. 123 In Tables 4.4.1 through 4.4.5 the relevant parameters for (a) Pre-cooked freeze-dried beef powder (10°C and 37.7°C) (b) Raw freeze-dried beef slices (10°C and 40°C) (c) Whole corn kernels (22°C and 50°C) (d) Wood cellulose (20°C and 60°C) and (e) Soda boiled cotton (10°C and 30°C), are shown. zn- in 0 tion possible for the product and temperature under consideration. The quantity E p2[ J in each table represents the best predic- Predicted isotherms based on these tables are represented in Figures 4.10 through 4.14. These plots together with the standard error estimates shown on each table are, for most of the points, within limits of the magnitude of error ordinarily associated with adsorp- tion experiments. The error estimates are also within the limits of variability to be expected on account of the highly Specific character of biological materials. The Corrected Density,pzi In rationalizing the corrected density term, p2, it should be kept in mind that the density function derived in this work, on which p1 is based, is acceptable only as a first estimate. In so far as the isosteric heat values determined by the combined usage of the Clausius-Clapeyron equation and the SIOpes of the so-called adsorption isostere plots are, at best, crude estimates of the actual quantities (Ross and Olivier, 1964; Gregg and Sing, 1967; Adamson, 1967), the ratio, (AH/AH°), of the derived density function is only as good as the isosteric heat values used. The relatively small correction factors needed to up-grade O1 to 02, in most cases, are justified on the basis of the associated isosteric heat difficiencies discussed above. Moreover, p2 is seen to conform 124 Table 4.4.1(a). Calculation of theoretical isotherm. Product Isotherm temperature = 10°C pre-cooked freeze~dried beef powder [211-110] ._ 0 0 Ma 9 ‘E 2 : 3.01 p SDZEZ ”x P/Po (Expt1.) 1 .454 m ' 2 Error .1 .070 1.360 " .0957 1.612 .070 .000 .2 .095 1.360 " .1450 1.451 .095 .000 .3 .110 1.316 " .1870 1.302 .110 .000 .4 .122 1.316 " .2290 1.160 .132 -.010 .5 .135 1.316 " .2750 1.160 .144 -.009 .6 .155 1.164 " .3260 1.160 .171 -.016 .7 .191 1.164 " .3900 1.160 .205 -.014 .8 .260 1.164 " .4790 1.160 .252 -.008 .9 .380 1.126 " .6350 1.160 .334 +.056 95 .450 1.062 " .7990 1.160 .420 +.030 .98 .483 1.00 " 1.0400 1.023 .483 .000 Standard Error Estimate = i.0064 125 Table 4.4.l(b). Calculation of theoretical isotherm. Product = pre-cooked freeze-dried beef powder Isotherm temperature = 37.7°C zn_xn E n J 7 .. 0 0 Ma 9 g '0. -= 0.2 50.,[2 -1 J P/Po (Expt1.) 1 81.26 PM ~ 0.1 02 Error .1 .0209 1.360 .0089 1.844 .020 +.009 .2 .0309 1.360 .0165 1.532 .031 .000 .3 .0391 1.316 .0240 1.292 .039 +.000 .4 .0500 1.316 .0319 1.248 .049 +.001 .5 .0591 1.316 .0407 1.248 .063 -.004 .6 .0791 1.164 .0512 1.248 .079 .000 .7 .1020 1.164 .0648 1.248 .100 +.002 .8 .1410 1.164 .0844 1.248 .132 +.009 .9 .2000 1.126 .1210 1.248 .190 +.01 .95 .2410 1.062 .1620 1.248 .254 -.013 .98 .2860 1.000 .2260 1.000 .284 +.002 Standard Error Estimate - i.0064 126 127 .50 1 T I I 1 I l I T A EXPERIMENTAL IO' 8 .454 .. 9 EXPERIMENTAL 37.720 .40L 5 .357 '1 3 3 .30~ a "2' THEORETICAL Is0TI-IERM 9 111 7) 8 0.1 I 5 P . 0I - I _ m , .. q 8 '25 6 . .454 [I m s / I- .20- A p n ‘3 I 0 I 2 A I '5”‘ I q 2. 0 - / A / .IO~ ._ ff 4 I 55/ THEORETICAL 2. [‘4 It ISOTHERM .(753’ ‘,a0" 17 a (3.2: " I” .r‘)’ F;.‘ Ctl ,I" E . l.383 .00 1 l 1 1 l 1 1 l .l .2 .3 .4 .5 .6 .7 .8 .9 LC P/P. Figure 4.10. Comparison of experimental adsorption isotherms with calculated isotherms for pre-cooked freeze-dried beef powder. Table 4.4.2(a). Calculation of theoretical isotherm. Product = raw freeze-dried beef slices (Data taken from Saravacos, 1965) Isotherm temperature 10°C. [Zn-1"] 11 E = n - 2n- 1” a p n = 0.01 §pz[ n J P/Po (Expt1.) 1 .1047 Pm = 0.10 92 Error .1 .030 1.360 " .214 1.382 .030 .000 .2 .060 1.360 " .390 1.382 .055 +.005 .3 .080 1.316 " .556 1.382 .080 .000 .4 .100 1.316 .724 1.351 .101 -.001 .5 .120 1.316 .906 1.294 .121 -.001 .6 .142 1.164 1.110 1.241 .143 -.001 .7 .170 1.164 1.370 1.204 .172 -.002 .8 .200 1.164 1.710 1.132 .202 -.002 .9 .250 1.126 2.280 1.064 .253 -.003 .95 ‘.310 1.062 2.860 1.050 .313 -.003 .98 .375 1.00 " 3.630 1.00 .38 -.005 Standard Error Estimate = 10.0027 128 Table 4.4.2(b). Calculation of theoretical isotherm. raw freeze-dried beef slices (Saravacos, 1965) Product = Isotherm temperature = 40°C M ‘E [2:Eflih (Theorztical) M 0 —. Zn_xn a = 0 - .04 =§Pz[ 0 J P/Po (Ethl') p1 .136 P111 3 '10 p2 Error .1 .027 1.36 .126 1.600 .027 .000 .2 .050 1.36 .230 1.600 .049 +.001 .3 .065 1.316 .329 1.520 .066 -.001 .4 .080 1.316 .430 1.380 .080 .000 .5 .090 1.316 .539 1.270 .092 -.002 .6 .100 1.164 .666 1.145 .103 -.003 .7 .120 1.164 .821 1.145 .127 -.007 .8 .155 1.164 1.030 1.145 .160 -.005 .9 .215 1.126 1.400 1.145 .217 -.002 .95 .250 1.062 1.770 1.076 .258 -.008 .98 .300 1.000 2.280 1.000 .310 -.010 Standard Error Estimate 129 = $0. 0048 40 l l 1 I I l | l I 9 EXPERIMENTAL IO° C } (Saravacos, .35 - A EXPERIMENTAL 40°C 1965) .30 e 4. A THEORETICAL ISOTHERM L - ——> "f 97 8 .0I I 3 pm . .l I ,_ '25 _ 6 = .I047 f” 2 / m g I‘ O .20 '- / ‘4 0 I In / 5 .l5 - f“ 4 53 / g xe‘ /’ _ ,A _ .IO ’4’ ,A"u-THEORETICAL . //A’ ISOTHE RM L . 4 1355 3’45 I" ,cyq '/ 8": J 6 - .I360 13c, 1 1 1 l J l I l J .l .2 .3 .4 .5 .6 .7 .8 .9 LG P/Po Figure 4.11. Comparison of experimental adsorption isotherms with calculated adsorption isotherms for raw freeze-dried beef in slices. 130 Table 4.4.3(a). Calculation of theoretical isotherm. Product = whole corn kernels (Chung and Pfost, 1967) Isotherm temperature = 22°C M g [7‘7“an a n - 0 x0 = n = -. 1 §pz[z .3 J P/PO (Expt1.) 91 .01415 Pm = .008 92 Error .1 .055 1.295 .0300 1.380 .058 -.003 .2 .072 1.295 .0436 1.215 .074 -.002 .3 .090 1.145 .0548 1.215 .094 -.004 .4 .105 1.145 .0654 1.188 .109 -.004 .5 .118 1.145 .0763 1.129 .121 -.003 .6 .132 1.145 .0882 1.104 .137 -.005 .7 .150 1.043 .1020 1.104 .159 -.004 .8 .182 1.043 .1200 1.104 .187 -.005 .9 .232 1.000 .1490 1.104 .232 0 .95 .266 1.000 .1790 1.081 .273 -.007 .98 .286 1.000 .2080 1.000 .294 -.008 Standard Error Estimate = £0.0046 131 Table 4.4.3(b). Calculation of theoretical isotherm. Product = whole corn kernels (Chung & Pfost, 1967) Isotherm temperature = 50°C M p " [211'an a 1 5 n ._ 27140 n =- - .08 502E771 P/Po (Expt1.) =.0184 Pm * .008 92 Error .1 .050 2.08 1.400 .053 -.003 .2 .068 3.04 1.240 .068 .000 .3 .080 3.83 1.146 .080 .000 .4 .092 4.58 1.106 .092 .000 .5 .105 5.36 1.106 .108 -.003 .6 .120 6.21 1.106 .126 -.006 .7 .141 7.22 1.106 .145 -.004 .8 .179 8.54 1.106 .173 +.006 .9 .230 10.70 1.106 .216 +.024 95 .260 12.60 1.106 .255 +.005 98 .287 15.20 1.000 .279 -.001 Standard Error Estimate = 10.0079 132 -30 I r I I r T I I I .23- 0 EXPERIMENTAL 22°C } (Chung & Pfost, 2 A EXPERIMENTAL 50°C 1967) I .26- p- .24— I _ "I .22— f — ., I 4120- / 4 g THEORETICAL ISOTHERM / F ,l 8r- ” "0.1 A .— a PMI .008 fl '5 .I8e 6 - .OI42 / a o ‘/ 6) MP /° — I: / :5 _ I *- .12 /A " Q o ’ 1’ 2 4'0 _. y/ —I A / .081- .. ,fi’ THEORETICAL ISOTHERM ‘ /°’ 1; . -.08 '06 - a / P” I .008 _ 6 - .OI84 .04 - I — .02 ’ .. l I I l I I l I I .I .2 .3 .4 .5 .6 .7 .8 .9 10 Figure 4.12. Comparison of experimental adsorption isotherms with calculated isotherms for corn (based on the empirical data of Chung and Pfost, 1967). 15') 9 v .5)... .l Table 4.4.4(a). Calculation of theoretical isotherm. Product = wood (Lewis, 1921) Isotherm temperature = 20°C .. [2";an 1“a E 1’1 ' - . 10 E D 2 [£212.93] P/PO (Expt1.) Pl .0172 PIn - 0.10 92 Error .1 .0250 1.153 1.33 1.22 .027 -.002 .2 .0495 1.153 2.40 1.22 .050 .000 .3 .0630 1.101 3.38 1.123 .065 -.002 .4 .0800 1.101 4.36 1.094 .081 -.001 .5 .0950 1.101 5.39 1.050 .097 -.002 .6 .1130 1.067 6.53 1.050 .116 -.003 .7 .1350 1.067 7.90 1.030 .139 -.004 .8 .1650 1.042 9.68 1.030 .171 -.006 .9 .2250 1.024 12.50 1.030 .221 +.004 .95 .2750 1.0065 15.10 1.030 .267 +.008 .98 .3080 1.000 18.30 1.000 .314 -.006 Standard Error Estimate 8 $0.00415 134 Calculation of theoretical isotherm. Product = wood (Lewis, 1921) Table 4.4.4(b). Isotherm temperature = 60°C M [221“] a E _ 2 1’1_ x“ n - -o.o7 ngE n J p/ro (Exptl.) 01 -.020 Pm - 0.10 Oz Error .1 .020 1.153 .783 1.249 .020 .000 .2 .035 1.153 1.410 1.224 .035 .000 .3 .045 1.101 2.000 1.145 .046 -.001 .4 .060 1.101 2.590 1.145 .060 .000 .5 .070 1.101 3.210 1.145 .075 -.005 .6 .089 1.067 3.910 1.145 .091 -.002 .7 .110 1.067 4.750 1.145 .111 -.001 .8 .140 1.042 5.870 1.145 .136 +.004 .9 .188 1.024 7.670 1.145 .178 +.010 .95 .225 1.0065 9.390 1.145 .218 +.oo7 .98 .246 1.000 11.600 1.050 .245 +.001 Standard Error Estimate = £0.00423 135 136 .32 I I I r I I I l I .30- A EXPERIMENTAL 20°C — 1 (Lewis, 1921) .23- 9 EXPERIMENTAL 80°C 5 26I— _. .24— f. 3.22 — 7 - '0' F_.2O~- / — z m ,8_ 9 _ F. . z / 8 16‘ A / _. ' *' THEORETICAL ISOTHERM / m . , _ g .14- 7) 0.1 A J _. ,_ Pu, . 0.! / g f - .OI72 / o .‘2 " I. " S ' / .IOP , fl/ 1 .08 P a / _ ,o’ . M/ THEORETICAL ~06 ' ’90’ ISOTHERM i “ ° lgr’ ‘0 "'-.(77 .o .— fl, Pm' mo _ 4 / - 0.02 .02 - ,6’ 5 ~ I l I J I I l I l .I .2 .3 4 .5 .6 .7 .8 .9 IO Figure 4.13. PIP. Comparison of experimental adsorption isotherms with calculated adsorption isotherms for wood cellulose (based on the data by Lewis, 1921). Table 4.4.5(a). Calculation of theoretical isotherm. Product = cotton (Urquhart and Williams, 1924) Isotherm temperature = 10°C M. [Zn-1n]~ ° 3' —~ zn—I" n . 0.10 €92E n J P/PO (Expt1.) p1 =.199 Pm . 0.01 92 Error .1 .023 1.19 .0957 1.243 .023 .000 .2 .032 1.19 .145 1.122 .032 .000 .3 .041 1.072 .187 1.122 .041 .000 .4 .049 1.072 .229 1.093 .049 .000 .5 .058 1.072 .275 . 1.078 .058 .000 .6 .068 1.045 .326 1.078 .070 -.002 .7 .080 1.045 .390 1.045 .081 -.001 .8 .099 1.045 .479 1.045 .099 .000 .9 .132 1.013 .635 1.043 .131 +.001 .95 .164 1.001 .799 1.032 .164 .000 .98 .209 1.000 1.04 1.014 .209 .000 Standard Error Estimate = 40.000738 137 Table 4.4.5(b). Calculation of theoretical isotherm. Product = cotton (Urquhart and Williams, 1924) Isotherm temperature = 30°C M ‘E [Zn-1n] Mé(Theoretical) a 0 _. 20-x” n - 0.20 §DZ['—Er-J P/Po (Expt1.) p1 =.799 PIn - 0.01 02 Error .1 .018 1.19 .0935 1.388 .018 .000 .2 .027 1.19 .141 1.311 .027 .000 .3 .035 1.072 .183 1.311 .035 .000 .4 .044 1.072 .224 1.311 .044 .000 .5 .052 1.072 .268 1.273 .052 .000 .6 .061 1.045 .319 1.273 .062 -.001 .7 .073 1.045 .381 1.221 .073 .000 .8 .091 1.045 .469 1.221 .092 -.001 .9 .121 1.013 .621 1.178 .121 .000 95 .147 1.001 .783 1.091 .147 .000 .98 .163 1.000 1.020 1.00 .185 .022 Standard Error Estimate - +0.0066 (based on 11 pts) 8 £0.000447 (based on 10 pts) 138 '24 T j T I I 1 I I I .224 9 EXPERIMENTAL IO°C 61 I (Urquhart and .20— A EXPERIMENTAL ao'c Williams, 1924) - 3 l . .Is- '. 3 I I- .IGP °A..I z I m ‘I . 5 .I4 e f‘ - o THEORETICAL ISOTHERM . , 0 17 - 0J0 .l2 - J B .10 I- E ' 0.197 o l A a ,5 O .08 +- 0 / _. 2 ’A/ ° / .06 - o ,{5 THEORETICAL - g I’A’ ISOTHERM -°4 _ z’ a; - 0.20 — 0’, . Pm- O.OI °°2 " , . 6 cor/99 " l l l l I 1 l 1 l ' .l .2 .3 .4 .5 .6 .7 .8 .9 LG P/ P, Figure 4.14. Comparison of experimental adsorption isotherms with calculated adsorption isotherms for soda-boiled cotton (data taken from Urquhart & Williams, 1924). 139 in every case, with the essential implications of the adsorption potential theory. In showing consistently increasing density values with rising isotherm temperatures, p2 reflects the correct tempera- ture dependence of the isosteric heat of adsorption as evident in the well known relation (Ross and Olivier, 1964): Ans: = qdiff + RgT where qdiff is the differential heat of adsorption. It appears reasonable, in the absence of conclusive empirical density values, to accept the O2 values on their demonstrated quali- tative merit. Empirical_g and E Parameters: A summary of the empirically deter- mined values of n and E for corn, cotton, wood, raw freeze-dried beef slices and pre-cooked freeze-dried beef powder is presented in Table 4.4.6. The values are clearly product and temperature dependent. However, when compared with their counterpart semi- theoretically determined n and E values (Table 4.1.3), drastic dif- ferences in magnitude and character are evident. These differences once again emphasize the fundamentally qualitative character of the pore-size distribution parameters, Y and K1, as determined. While temperature induced-changes in the pore-structure were generally predictable from the figures of Table 4.1.3, certain interpretations deduced from those figures still need to be revised or finalized. Since the empirical values of n and E (Table 4.4.6) show increasing trends from smaller to larger values with rising isotherm temperatures, a progression from smaller to larger pores 140 Table 4.4.6. Summary of empirical n, and E parameters. PRODUCT Temperature 0 P111 5 °c 22.0 -0.10 .008 .01415 CORN 50.0 - .08 .008 .01840 10.0 0.10 0.01 .19900 COTTON 0.20 0.01 .79900 20.0 - .10 0.10 .01720 woos " a 07 0. 10 a 02000 RAW 10.0 0.01 0.10 0.10470 FREEZE-DRI ED BEEF SLICES 40.0 0.04 0.10 0.13600 FREEZE-DRIED BEEF ’°”DER 37.7 0.2 0.10 1.26000 141 is indicated. .This expansion in the pores must be accompanied by a concomitant contraction in the solid component of the porous structure. From tabulated Pm values (Table 4.4.6), it is at once obvious that the Pm parameter as used in this study is quite different from the actual pressure corresponding to the mono-layer capacity. Part of this discrepancy can be traced back to the transformation of coordinate axes which became necessary in the original formulation in chapter II. It would appear that in the attempt to systematize the plots of the derived isotherm equation, the actual physical significance of Pm has been sacrificed. Stability Considerations in the Sorbent Pore Structure: The standard error estimate of i 0.021 calculated in Table 4.4.1 8 for pre- cooked freeze-dried beef powder is by far the worst recorded. Since this product happens to be the only biological material in its pow- dered form tested, the comparatively poor error estimate raises a question of structural stability in powdered adsorbents. It would seem that the pore structure of a powdered adsorbent is basically ill defined. For one thing, the pore sizes and pore- size distributions associated with such sorbents must necessarily change with the degree of sorbent compaction. It is also clear, that both hydro and thermo induced changes in the sorbent pore structure will be more pronounced during the adsorption process. These trends appear to be well reflected not only by the compara- tively poor error estimate of Table 4.4.1 a, but also by (a) the change from D - 0.1 and 5 - .454 at 10°C (Table 4.4.1 8) to 142 ‘0 = 0.2 and E = 1.26 at 37.7°C (Table 4.4.1 b), and (b) the accom- panying change in Pm value from 0.01 at 10°C to 0.1 at 37.7°C. Closure: It has been demonstrated in this section that the derived isotherm equation can be used to reconstruct the adsorption isotherms of biological materials with a fair degree Of accuracy. Even though certain quantitative inadequacies of present methods made impossible a priori determination of the structural parameters, 0 and E, it is still possible to formalize the temperature dependence of these parameters either on a general or specific basis. While no such attempt has been made in the present study because of the limited number of isotherm temperatures investigated, the necessary founda— tions appear to have been laid to make such an effort a logical extension of the present work. 143 4.5 Verification of the PrOposed "Capillary Condensation - SwellingfiFatigue" Doublg Superposition Theory of Sorption Hysteresis in Biological Materials As an example, the estimation of desorption isotherms from adsorption data, employing the scheme outlined in section 2.5 c, is implemented in Tables 4.5.1 and 4.5.2 for corn kernels at isotherm temperatures of 22°C and 50°C. Corn is used because, to the knowledge of the author, it is the only bio-material of which the bulk moduli have been empirically determined for a wide range of moisture con- tents (White, 1966). The work sheets (Tables 4.5.1 and 4.5.2) are deveIOped as follows: column 2 represents calculated or experimental adsorption data; in this case the data of Chung and Pfost (1967) for whole corn kernels were used. Column 3 gives the number of adsorbed multi- layers, n = X/Xm. Columns 4 and 5 are the approximate density and volumetric data. The terms of column 6 are obtained by linear inter— polation Of the incremental volumes; in this case each incremental volume is divided by a factor of 10- to give suitably small adsorp- tion steps. Column 7 gives the bulk moduli data Obtained by extra- polation of the data by White (1966). Columns 8 and 9 are calculated in accordance with equations (2.5.17a) and (2.5.18), respectively. Columns 10, 11 and 12 Show trial adsorption relative vapor pressure values Obtained by the selective variation of the parameter, w, in the generalized Cohan relation (eqn.(2.5.4)). In this case, m was found to have the value 1.25. Thus the desorption relative vapor pressure values calculated with the equation: 144 x no z.o o Na> Ao\omo ca 9 m 5.5A n so 4 145 oo.a oo.a oo.H oooN. ooo. oo.o . oom. ooo.a ooN. o.” who. do». How. Anna. Haoo. on.e omo. Nmm. ooo.a Now. o. we». was. was. mood. naoo. oo.m omo. as“. neo.a and. o. owe. «me. who. oNnH. ooNoo. om.o amo. mes. moo.a on“. a. can. «on. an. mama. naoo. oa.a oao. maa. nea.a «ma. 6. cos. nae. ooe. ooua. ouoo. an.“ ago. nod. nee.” wag. m. «on. mom. «on. smog. aeoo. ma.» Mao. ooo. noa.a mod. 4. oou. com. com oooo. oooo. oo.o mac. who. med.“ ooo. n. «ma. no”. AH. Noao. Nooo. om.o mNo. moo. mo~._ mao. N. mo. moo. mo. +ommo. nao «so. nou.a mno. a. Neo. Aom\omv ozo+mzu AoooHoMHszo Aoa xv c A.auome m.aaom\mo ~.HAom\mv H.~Aom\ao 62 one one a NH. 6> a e 6: om\m NH AA oa o o 4 a o m 4 m a a .oomm u eunumwooEmu EomsuomH Anomfi .umomm a mangov Hmcumx once u uosooum .eConuomH cowomuomoo mo coauoaooaoo one cow ooozm xuoz .H.m.q mapmw x no x a o m u so 4 Nao Aa\oooca e a n.4H oooo.a ooom. oooo. oo.e ooom. ooo.a o om. o.a oooo. moan. Nana. Naoo. om.o oomu. ooo.a ma. o. oomo. m ooma. oooa. oaoo. om.n oaaa. Mao.“ oaa. o. oano. ooqo. owed. muoo. om.o coma. moo.a moo”. a. cane. N ooNn. «ANA. naoo. oa.a oooa. msa.a NA. o. . omao. ,6 come. mooa. oaoo. nm.~ oaoo. mea.a mod. n. “m Roao. . ooan. oooo. oeoo. ma.» mono. mea.a «mo. 4. none. oNNN. omoo. onoo. oo.o oooo. noa.a oo. n. naao. a onma. oeao. oooo. om.o ammo. moN.H woo. N. omao. oono. oomo. oomo. mow.” omo. a. oomo. Aoooa.ooapzv Aooaxo Aoaxo a A.Auoxav m~.HAom\mo u oo\oa Aozc+mzv 4624 x «Ho mo 6 e 6: om\a oa o o a o m e n N A Doom u musumuanou EnenuomH Amooa .umme a wcssuv Hmcumx upon I uoswoum .ECoLooma coHuauomoo we cowowaooamo one new noose xuos .N.m.¢ dance _ 1.25 Pd/Po — (Pa/Po) [4.5.1] were plotted against the corresponding specific desorbed mass Md (column 9) to obtain the theoretical desorption curves (Figure 4.15). Figure 4.15 Shows that the theoretical desorption curves fit the experimental points remarkably well for the isotherm temperatures investigated. While broad generalizations must await more extensive application of the theory to other biological products and other iso- therm temperatures, this finding indicates that the new approach pre- sented here offers an initial framedwork for more detailed considera- tion of the complex questions of hysteresis in bio-materials. fi_._5__§. The parameter, _u)_, _a_s_ 3 factor of pore geometry A convenient classification of sorption hysteresis based on (a) the pressure range over which the main boundary 100p extends, and (b) the steepness of the adsorption and desorption branches, has been prOposed by de Boer (1958). Within the frame-work of this broad classification, the parameter, m, can be employed to gain more insight into the geometrical shape of the pores of sorbing biological materials (Barrer st 31., 1956; de Boer, 1958; Linsen and Van Den Heuvel, 1967). Thus, for the corn kernel, equation (4.5.1) satisfies the de Boer (1958) criterion: (Pa/Po)2 < Pd/Po [4.5.21 Consequently, the pores are tubular capillaries with widened spheroidal parts Open at both ends. This finding supports the earlier preference in section 4.1 for the interconnected spheroidal ink bottle model. 147 .30 T I 0 EXPERIMENTAL 22.0 A EXPERIMENTAL 50°C 0) (I I 09 C) I THEORETICAL L. ISOTHERM (I A’ _. I .IA’ ‘\r THEORETICAL ISOTHERM MOISTURE CONTENT (d. b.) 5 .05 .I .2 .3 .4 .5 .6 .7 .8 .9 IO P/Po Figure 4.15. Comparison of experimental desorption isotherms with calculated desorption isotherms for corn (expl. data taken from Chung and Pfost, 1967). 148 V. SUMMARY AND CONCLUSIONS A generalized model of water sorption is developed and verified for biological materials. The B-E-T and Capillary condensation theories are combined into an integral isotherm equation for porous biological materials. In order to solve the integral equation explicitly, it became necessary to: (a) characterize analytically, the physical structure of sorbing (b) bio-materials in terms of a pore-size distribution function and an idealized pore-geometry; derive, utilizing the general framework of the potential theory in conjunction with the fundamental laws of thermo- dynamics, an explicit density function for water adsorbed by biological materials. As an essential complement for this method of attack, a new theory of sorption hysteresis is syn- thesized within the generalized framework. In consequence of a detailed application of the deve10ped model to the sorption data of several biological products, the following specific conclusions can be drawn. 1. The pores of biological materials are best modeled for trans- port processes in these systems as interconnected spheroidal "ink-bottles". The numerical distribution of pore-Size in bio-materials is well described by a power law distribution function of the type: _ Y m2(R) — KlR 149 where the pore characterization parameters, K and Y are 1 both product and temperature dependent. The density of water sorbed by biological materials can be expressed, in the first approximation, as the semi-theoretical relation: p/po = 0* [Mist/Mist:l where p is the density associated with the isosteric heat of adsorption, AH po and Afizt are the corresponding density st’ and isosteric heat values at saturation vapor pressure; and the quantity, u*, is empirically defined for the class of biological materials of interest. The interconnected spheroidal ink-bottle pore model, the power law distribution function and the density function are integrated into an isotherm equation of the form: a M = 9715 [zn- in] where n and g are the primary and secondary characteristic para- meters of pore structure; and Z and 1 are functions of relative vapor pressure. The derived isotherm equation, (a) reflects through its density term, the well established fact that the heat of adsorption varies with moisture content, (b) reconstructs the empirical adsorption isotherm and its dependence on temperature reasonably well, (c) yields most characteristic isotherm types principally through the variation of the principal structural parameter, 0. 150 It thus appears conclusive that the three basic concepts in adsorption, namely: (a) the Kinetic concept as exemplified by the B-E-T theory of multi-molecular adsorption, (b) Polanyi's adsorption potential theory, and (c) Zsigmondy's capillary con- densation theory, play complementary and possibly overlapping roles and can be unified into a coordinated theory of water sorption by biological materials. A new theory of sorption hysteresis based on the superposition of (a) Capillary Condensation hysteresis, and (b) Swelling Fatigue hysteresis, is developed in quantitative terms and has been successfully employed to predict the desorption isotherms of whole corn kernels. 151 VI. SUGGESTIONS FOR FURTHER STUDY As a consequence of the present study, it is recommended that further investigations should be undertaken in the following areas: 1. W. The function derived in this study for the density of water can be improved considerably if (a) more precise isosteric heat values are measured and (b) the determina- tive function, u*, is revised to reflect a wider spectrum of biological materials. Furthermore, because of the almost complete absence of reliable experimental data on the density of water sorbed on biological materials, the derived density function is yet to be conclusively verified. The Structural Parameters, n and g. The dependence of n and g on temperature needs to be formalized in order to make it possible to extrapolate from a set of empirical TI and g values at one isotherm temperature to another. If these parameters are deter- mined for a sufficiently large number of biological materials (over a wide enough temperature range) formalization on a general basis would be possible. Furthermore, water-induced Changes in biological materials should be studied in terms of the 0 and 5 parameters. In this case the structural parameters determined from a desorption isotherm can be compared with those resulting from the adsorption branch. This type of parametric study can conceivably furnish additional information on the phenomenon of hysteresis. 152 39 Mechanical PrOperties of Biological Materials. The bulk modulus of most biological materials as a function of moisture content and temperature has not been experimentally determined. Therefore, it was not possible to test the "Capillary Condensa- tion - Swelling Fatigue" superposition concept of sorption hysteresis developed here in sufficient detail to establish its general validity. Additional work should be done along these lines. 153 REFERENCES Adamson, A. W. (1967). The Physical Chemistry of Surfaces. (second edition), Interscience Publishers, New York. Alfrey, T. and P. Doty (1945). The Methods of Specifying the PrOperties of Viscoelastic Materials. J. of Appl. Physics. 16:700. Allamand, A. J., P. G. T. Hand and J. E. Manning (1929). The Sorption of Water by Activated Charcoals, Part V. J. of Physical Chemistry. 33:1694. Bakker-Arkema, F. W. (1961). Desorption and Adsorption Isotherms for Alfalfa of Four Growth Stages in the Temperature Range of 40° to 120°F. Unpublished M.S. Thesis, Agricultural Engineering Department, Michigan State University. Barrett, E. P., L. G. Joyner and P. P. Halenda (1951). The Determination of Pore Volume and Area Distributions in Porous Substances. J. Am. Chem. Soc. 73:373. Babbitt, J. D. (1942). On the Adsorption of water Vapor by Cellulose. Canadian J. of Research. 20(A), No. 9, pp. 143-172. Bangham, D. H. and W. Sever (1925). An Experimental Investigation of the Dynamical Equation of the Process of Gas-Sorption. Phil. Mag. 6(49):93S. Barkas, W. W. (1953). Part A,AMechanical PrOperties of Wood and Paper. R. Meredith, Editor, Interscience Publishers, New York. Barrer, R. M., N. McKenzie and J. S. S. Reay (1956). Capillary Condensation in Single Pores. J. of Colloid Science. 11:479. Becker, H. A. and H. R. A. Sallans (1956). A Study of Desorption Isotherms of Wheat at 25°C and 50°C. Cereal Chem. 33:79. Bernal, J. D. and R. H. Fowler (1933A). A Theory of Water and Ionic Solutions, with Particular Reference to Hydrogen and Hydroxyl ions. J, Chem. Physics. 1:515. Bernal, J. D. and R. H. Fowler (1933B). Pseudocrystalline Structure of Water. Transactions Faraday Soc. 29:1049. Benson, S. W., D. A. Ellis and R. W. Zwanzig (1950). Surface Area of Proteins III. Adsorption of Water. J. Am. Chem. Soc. 72:2102. Benson, S. W. and J. W. King (1965). Electrostatic Aspects of Physical Adsorption: Implications for Molecular Sieves and Gaseous Anesthesia. Science. 150:1710. 154 Bewig, K. W. and W. A. Zisman (1964). Surface Potentials and IndUCed Polarization in Nonpolar Liquids Adsorbed on Metals. J. Physical Chemistry. 68:1804. Bradley, R. S. (1936). Polymolecular Adsorbed Films. Part I. The Adsorption of Argon on Salt Crystals at Low Temperatures and the Determination of Surface Fields. J. Chemical Society. 1467-1474. Brunauer, S., P. H. Emmett and E. Teller (1938). Adsorption of Cases in Multi-molecular Layers. J. Am. Chem. Soc. 60:309. Brunauer, 8., L. S. Deming, W. E. Deming and E. Teller (1940). On a Theory of Van Der waal's Adsorption of Gases. J. Am. Chem. Soc. 62:1723. Brunauer, S. (1945). The Adsorption of Gases and Vapors Vol. 1. Princeton University Press, Princeton. Brunauer, 8., L. E. Capeland and D. L. Kantro (1967). The Langmuir and B-E-T Theories. In Solid-Gas Interfagg_Vol. 1, p. 77. Edited by Flood. Marcel Dekker Inc., New York. Burrage, L. J. (1934). Studies on Adsorption. Part VII. The Form of Isothermals of Vapor on Charcoal and Its Relation to Hysteresis. Trans. Faraday Soc. 30:317. Cassel, H. M. (1944). Cluster Formation and Phase Transitions in the Adsorbed State. J. Phys. Chem. 48:195. ""Chung, D. S. and H. B. Pfost (1967). Adsorption and DeSorption of Water Vapor by Cereal Grains and Their Products. Parts I, II and III. Transactions A.S.A.E. 10(4):552. Clampitt, B. H. and D. E. German (1958). Heat of Vaporization of Molecules at Liquid-Vapor Interfaces. J. Phys. Chem. 62:438. Coelingh, M. B. (1938). Thesis, University of Utretch. Coelingh, M. B. (1939). Optische Untersuchungen fiber das FlUssigkeit Dampfgleichgewicht in Kapillaren Systemen. Kolloid Z. 87:25]. Cohan, L. H. (1938). Sorption Hysteresis and the Vapor Pressure of Concave Surfaces. J. Am. Chem. Soc. 60:433. Cohan, L. H. (1944). Hysteresis and the Capillary Theory of Adsorption of Vapors. J. Am. Chem. Soc. 66:98. Coulson, C. A. (1959). Hydrogen Bonding. Edited by D. Hadzi. Perganon Press Ltd., London. 155 Cranston, R. W. and F. A. Inkley (1957). The Determination of Pore Structures from Nitrogen Adsorption Isotherms. Advances in Catalysis. 9:143. Davidson, G. F. (1927). The Specific Volume of Cotton Cellulose. J. Textile Institute. 18:T 175. Day, D. L. and G. L. Nelson (1965). Desorption Isotherms for Wheat. Trans. A.S.A.E. 8(2):293. de Boer, J. H. and C. Zwikker (1929). Adsorption als Folge von Polarisation die Adsorptionsisotherme. Z. Physik Chem. 83:407. de Boer, J. H. (1953). The Dynamical Character of Adsorption. Clarendon Press, Oxford. de Boer, J. H. (1958). The Shapes of Capillaries. In Structure and Properties of Porous Materials. Edited by Everett and Stone. Butterworths Scientific Publications, London. p. 68. Defay, R., I. Prigogine, A. Bellemans and D. H. Everett (1966). Surface Tension and Adsorption. Wiley and Sons, Inc. New York. Dellyes, R. (1963). Modification A La Théorie B-E-T Et Nouvelles Possibilities D' Application. J. Chim. Phys. 60:1008. De Vries, Thos. (1935). Densities of Adsorbed Cases 1. Carbon Dioxide on Charcoal. J. Am. Chem. Soc. 57:1771. Dollimore, D. and G. R. Heal (1964). An Improved Method for the Calculation of Pore-Size Distribution from Adsorption Data. J. of Appl. Chem. pp. 109-114. Dorsey, N. E. (1940). PrOperties of Ordinary Water Substance. Reinhold Publishing Co., New York. Dubinin, M. M. (1955). A Study of the Porous Structure of Active Carbons Using a Variety of Methods. Quarterly Review (London). 9:101. Dubinin, M. M. (1965). Theory of Bulk Saturation of MicrOporous Activated Charcoals During Adsorption of Gases and Vapours. Russian J. of Phys. Chem (English Trans.). 39:697. Everett, D. H. (1958). Some Problems in the Investigation of Porosity by Adsorption Methods. In Structure and Properties of Porous Materials. Edited by Everett and Stone. Buther- worths Scientific Publications, London. p. 95. Everett, D. H. (1967). Adsorption Hysteresis. In The Solid—Gas Interface. Edited by Flood. Marcel Dekker, Inc. New York. p. 1055. 156 Ewing, A. (1881). Proc. Royal Society (London), 33:22. Ewing, D. T. and C. H. Spurway (1930). The Density of water Adsorbed on Silica Gel. J. Am. Chem-Soc. 52:4635. Filby, E. and 0. Maass (1932). The Volume Relations of the System, Cellulose and Water. Can. J. of Research. 7(2):l62. Flood, E. A. (1967). The Solid-Gas Interface Vols. I & II. Marcel Dekker, Inc., New York. Foster, A. G. (1932). The Sorption of Condensable Vapors by Porous Solids. Part I. The Applicability of the Capillary Theory. Trans. Faraday Soc. 28:645. Foster, A. G. (1934). The Sorption of Methyl Alcohol by Silica Gels. Proc. ngal Soc. (Londo_l. A. 146:129. Foster, A. G. (1948). Pore Size and Pore Distribution. Discus- sions Faraday Soc. 3:41. Foster, A. G. (1951). Sorption Hysteresis I. Some Factors Deter- mining the Size of the Hysteresis LOOp. Journal of Physical and Colloid Chemistgy. 55:638. Frenkel, Y. I. (1946). Kinetic Theory of Ligpids. The Clarendon Press, Oxford. Reprinted by Dover Publications, 1955. Gibbson, R. E. (1934). The Influence of Concentration on the Com- pression of Aqueous Solutions of Certain Sulfates and a Note on the Representation of the Compression of the Aqueous Solution as a Function of Pressure. J. Am. Chem. Soc. 56:4. Gorter, C. J. and H. P. R. Frederikse (1949). A Few Remarks on Physical Adsorption. Physica. 15:891. Gregg, S. J. and K. W. Sing (1967). Adsorption Surface Area and Porosity. Academic Press, London and New York. Guggenheim, E. A. (1967). Thermodynamics. North Holland Publishing Co., Amsterdam. Halsey Jr., C. D. (1948). Physical Adsorption on Non-Uniform Sur- faces. Journal of Chemical Physics. 16:931. Halsey Jr., C. D. (1950). The Role of Heterogeneity in Adsorption and Catalysis. Discussions Faradgy Soc. 8:54. A Hammerle, J. R. (1968). Failure in a Thin Viscoelastic Slab Sub- jected to Temperature and Moisture Gradients. Unpublished Ph.D. Thesis Penn. State University. 157 Harkins, W. D. and D. T. Ewing (1921). A High Pressure Due to Adsorption, and the Density and Volume Relations of Charcoal. The Compression of Liquids by Charcoal. J. Am. Chem. Soc. 43:1787. Harkins, W. C. and G. Jura (1944). Surfaces of Solids XIII. A Vapor Adsorption Method for the Determination of Area of Solids Without Assumption of a Molecular Area and Areas Occupied by Nitrogen and other Molecules on the Surface of a Solid. J. Am. Chem. Soc. 66:1366. Harvey, E. N. (1943). Surface Areas of Porous Materials Calculated from Capillary Radii. J. Am. Chem. Soc. 65:2343. Henderson, 8. M. (1952). A Basic Concept of Equilibrium Moisture. Agricultural Engineering 33(1):29. Henderson, S. M. (1969). Equilibrium Moisture Content of Small Grain Hysteresis. A.S.A.E. paper No. 69-329 presented at Purdue University, June 22-25, 1969. Hill, T. L. (1952). Theory of Physical Adsorption. Advances in Catalysis. 4:211. Hill, T. L. (1960). Introduction to Statistical Thermodynamics. Addison-Wealey Publishing Co., Inc., Reading, Mass. and London. Hfittig, G. F. (1948). Zur Auswertung der Adsorptions-Isothermen. Monatsh. Chem. 78:177. Jastrzebski, Z. D. (1957). Nature and PrOperties of Engineering Materials. Wiley and Sons, Inc., New York. Jungh (1965). Ann Phys. Chem. 125:292. Katz, S. M. (1933). The Laws of Swelling. Transactions Faraday Soc. 29:279. Katz, S. M. (1949). Permanent Hysteresis in Physical Adsorption. J. Phys. Chem. 53:1166. Kraemer, E. 0. (1931). A Treatise on Physical Chemistry. Edited by H. S. Taylor. D. Van Nostrand Co., New York. Kuprianoff, J. (1958). Bound Water in Foods. In Fundamental Aspects of Dehydration of Foodstuffs. London, England: Society of Chemical Industries. KUhn, I. (1964). A New Theoretical Analysis of Adsorption Phenomena. Introductory Part: The Characteristic Expression of the Main Regular Types of Adsorption Isotherms by a Single Simple Equation. J. Colloid Science. 19:685. 158 Labuza, T. P. (1968). Sorption Phenomena in Foods. Food Technology. 22:263. Lamb, A. B. and A. S. Coolidge (1920). The Heat of Adsorption of Vapors on Charcoal. J. Am. Chem. Soc. 42:1146. Lambert, B. and A. G. Foster (1932). Studies of Gas-Solid Equili- bria. Part IV. Pressure-Concentration Equilibria between Ferric Oxide Gels and (a) Water, (b) Ethyl Alcohol, (c) Benzene, directly Determined under Isothermal Conditions. Proc. Roy. Soc. (London) A136, 363. Langmuir, I. (1918). The Adsorption of Cases on Plane Surfaces of Glass and Mica and Platinum. J. Am. Chem. Soc. 40:1361. Lewis, W. K. (1921). The Rate of Drying of Solid Materials. The Journal of Industrial and Engineering Chemistry. Lippens, B. C., B. G. Linsen and J. H. de Boer (1964). Studies on Pore Systems in Catalysis. I The Adsorption of Nitrogen; Apparatus and Calculation. II The Shapes of Pores in Aluminum Oxide Systems. 111 Pore-Size Distribution Curves in Aluminum Oxide Systems. J. of Catalysis. 3:32. Lisen, B. G. and A. Van den Heuvel (1967). Pore Structure. In The Solid-Gas Interface. Edited by Flood. Marcel Dekker, Inc. New York. p. 1025. Lowry, H. H. and P. S. Olmsted (1927). The Adsorption of Cases by Solids with Special Reference to the Adsorption of Carbon Dioxide by Charcoal. J. Phys. Chem. 31:1601. Lykov, A. W. (1955). Experimentelle und Theoretische Grunlagen der Trocknung. V.E.B. Verlag Technik. Berlin, Germany. Matz, S. A. (1965). Water in Foods. A.V.I. Publishing Co., Inc., Westport, Connecticut. McBain, J. W. (1935). An explanation of Hysteresis in the Hydration and Dehydration of Gels. J. Am. Chem. Soc. 57:699. McMillan, W. G. and E. Teller (1951). The Role of Surface Tension in Multi-layer Gas Adsorption. J. Chem. Phys. 19:25. McMillan, W. G. and E. Teller (1951). The Assumptions of the B-E-T Theory. J. Phys. Chem. 55:17. McGavack, J. and W. A. Patrick (1920). The Adsorption of Sulfur Dioxide by the Gel of Silicic Acid. J. Am. Chem. Soc. 42:946. 159 Mfiller, H. (1882). On the Relation between Moisture Present in Textile Fabrics and that in the Atmosphere. J. Soc. of Chem. Industries. 1:356. Parks, G. J. (1902). On the Heat Evolved or Absorbed when a Liquid is Brought in Contact with a Finely Divided Solid. Phil. Mag. 4:240. Patrick, W. A. and F. V. Grimm (1921). Heat of Wetting of Silica Gel. J. Am. Chem. Soc. 43:2144. Patrick, W. A. (1929). Colloid Symposium Annual. 7:129. Pichler, H. J. (1956). SorptioniSOthermen fifir Getreide und Raps. (Sorption isotherms of wheat and rape-seed). Landtechnische Forschung 2:47. Pickett, G. (1945). Modification of the B-E-T Theory of Multi- molecular Adsorption. J. Am. Chem. Soc. 67:1958. Pierce, C. and R. N. Smith (1950). Adsorption-Desorption Hysteresis in Relation to Capillary of Adsorbents. J. Phys. Chem. 54:784. Pierce, C. and R. N. Smith (1950). Heats of Adsorption IV. Entrepy Changes in Adsorption. J. Phys. Chem. 54:795. Pierce, C. (1953). Computation of Pore Sizes from Physical Adsorption Data. J. Phys. Chem. 57:149. Pierce, F. T. (1929). A Two-Phase Theory of the Absorption of Water Vapor by Cotton Cellulose. J. Textile Inst. 20:133. Polanyi, M. (1914). Verhandl. Deutsch. Physich. Ges. 16:1012. Polanyi, M. (1916). Verhandl. Deut. Physik. Ges. 18:55. Polanyi, M. (1920). Z. Elektro. Chgg. 26:370. Rao, K. S. (1941). Hysteresis in Sorption. I, II, III, IV, V, VI. J. Phys. Chem. 45:501. Rodriquez-Arias, J. H. (1956). Desorption Isotherms and Drying Rates of Shelled Corn in the Temperature Range of 40 to 140°F. Unpublished Ph.D. Thesis, Dept. of Agri. Eng., Michigan State University. Rose (1849). Ann. Phys. 73:1. Ross, 8. and J. P. Olivier (1964). On Physical Adsorption. Interscience Publishers, New York, London, Sidney. 160 Saravacos, G. D. and R. M. Stinchfield (1965). Effect of Temperature and Pressure on the Sorption of Water Vapor by Freeze-Dried Food Materials. J. of Food Science. 30(5):779. Shull, C. G. (1948). The Determination of Pore-Size Distribution from Gas Adsorption Data. J. Am. Chem. Soc. 70:1405. Stamm, A. J. and R. M. SeBorg (1935). Adsorption Compression on Cellulose and Wood 1. J. Phys. Chem. 39:133. Stamm, A. J. and W. K. Loughborough (1935). Thermodynamics of Swelling of Wood. J. Phys. Chem. 39:121. Stamm, A. J. and L. A. Hansen (1937). The Bonding Force of Cellulose Materials for Water (from Specific Volume and Thermal Data). J. Phys. Chem. 41:1007. Stamm, A. J. (1938). Calculations of Void Volume in Wood. Indus- trial and Engr. Chem. 30(11):1280. Strohman, R. D. and R. R. Yoerger (1967). A New Equilibrium Mois- ture-Content Equation. Trans. A.S.A.E. 10(5):675. Swan, E. and A. R. Urquhart (1927). Adsorption Equations - A Review of the Literature. J. Phys. Chem. 31:251. Ter Haar, D. (1950). Phenomenological Theory of Visco-Elastic Behavior. Physica. 16:719, 738, 839. Thompson, T. L., R. M. Peart and G. H. Foster (1967). Mathematical Simulation of Corn Drying - A New Model. A.S.A.E. Paper No. 67-313. Presented at Saskatoon, Saskatchewan, Canada, June 27-30, 1967. Thomson (Lord Kelvin), W. (1871). On the Equilibrium of Vapor at a Curved Surface of Liquid. Phil. Mag. S.4 Vol. 42, No. 282: 448. Triebold, H. O. and W. A. Aurand (1963). Food Composition and Analysis. D. Van Nostrand Co., Inc. Princeton, New Jersey. Tuck, N. G. M., R. L. McIntosh and O. Maass. The Density of Adsor- bates. Can. J. of Research. 268:21. Urquhart, A. R. and A. M. Williams (1924). The Moisture Relations of Cotton. The Effect of Temperature on the Absorption of Water by Soda-Boiled Cotton. J. Textile Institute. 21:T559. Urquhart, A. R. (1929). Adsorption Hysteresis. J. Textile Inst. 20:T117. 161 Urquhart, A. R. (1929). The Mechanism of the Adsorption of Water by Cotton. J. Textile Inst. 20:125. Urquhart, A. R. and N. Eckersall (1930). The Moisture Relations of Cotton. VII. A Study of Hysteresis. J. Textile Inst. 21:T499. Van Laar (1924). Zustandsgleichung von Gasen und Flussigkeiten. Quoted by Lowry and Olmsted (1927), J. Phys. Chem. 31:1601. ”VEHISCIEI, w. B. and M. J. Copley (1963), FOOd Dehydration. Vol. 1. A.V.I. Publishing Co., Westport, Connecticut. Viswanathan, B. and M. V. C. (1967). Computation of Pore Size Distribution in Terms of Surface Area. J. of Catalysis. 8:312. Ward, A. G. (1962). The Nature of the Forces Between Water and the Micromolecular Constitutents of Food. In Recent Advances in Food Science. Vol. 3:207. Edited by Keitch and Rhodes. Butterworth's London. Wheeler, A. (1945). Report #8-9829. Circulated to the PAW ”Recom- mendation 41 Group" of the Petroleum Industry, June 1945; presented at A.A.A.S. Gordon Conference on Catalysis 1945 and 1946. Wheeler, A. (1955). Reaction Rates and Selectivity in Catalyst Pores. In Catalysis - Fundamental Principles. Vol. II, p. 105. Edited by P. H. Emmett. White, H. J. and H. Eyring (1947). The Adsorption of Water by Swelling High Polymeric Materials. Textile Research Journal. 17(10):523. White, R. K. (1966). Swelling Stress in Corn Kernel as Influenced by Moisture Sorption. M.S. Thesis in Agricultural Engineering Penn. State University. Young, D. M. and A. D. Crowell (1962). Physical Adsorption of Gases, Butterworth, London. Young, J. H. and G. L. Nelson (1967). Research of Hysteresis Between Sorption and Desorption Isotherms of Wheat. Trans. A.S.A.E. 10(6):756. Young, J. H. and G. L. Nelson (1967). Theory of Hysteresis Between Sorption and Desorption Isotherms in Biological Materials. Trans. A.S.A.E. 10(2):260. Zsigmondy, R. (1911). Z. Anogg. Chem. 71:356. 162 APPENDICES 163 APPENDIX A Isosteric heat of adsorption plotted as a function of moisture content for several materials. 164 AH” - (CAL./GRAM) 800 750 700 650 600 550 500 ?\ o SODA- 3011.50 COTTON X (from Urquhart and Williams, 1924) \ A SPRUCE WOOD \ (from Stamm, 1938) > El CORN KERNELS (DISSPT'N). \\ (from Rodriquez-Arias, 1956) \x 0 CORN ADEPTN \. (Chung & Pfost, 1967) \ \ (Chung & Pfost, 1967) llillLLLllllllllll11111111111 V 0.00 .04 .08 .I2 .16 .20 .24 .28 X-MOISTURE CONTENT (d.b.) Figure A-l. The isosteric heat of adsorption as a function of moisture content. 165 M., - (CAL. IGRAM) \A\ 0 FIRE-COOKED FREEZE-DRIED BEEF POWDER (Adptn.) \\ A Pas-000K150 sneeze-omen sees pownm wept...) 900— soo~ 700— 600— 550 I : - 0.00 0.10 0.20 0.30 0.40 0.50 0.60 X - MOISTURE CONTENT (d.b.) Figure A-2. The isosteric heat of sorption as a function of moisture content. 166 APPENDIX B Deve10pment of the working equation of the Cranston and Inkley (1957) method for the calculation of pore-size distribution from sorption isotherms. Let Vrér be the volume of pores having radii between r and r + 0r, where 6r is small compared with r. Consider an adsorption step from a relative pressure Pr such that the small pore in the range is about to fill with condensate, to a pressure P ), such (r+0r that the largest pore in the range has just filled with condensate. During this pressure change, pores in the range considered became filled with condensate, smaller pores are already filled, while in larger pores the thickness of the adsorbed layer on their walls in- creases from tr to tr + 6t. The total volume of liquid water adsorbed in this step is given by the volume of liquid condensed in pores with average radius '? 8 (r1 + r2)/2, and the volume of water con- densed in larger pores contributing to the thickness of the adsorbed layer. Thus, assuming a cylindrical geometry, 2 v_r.5r ”(r - tr) Lt + c(tr - tfiér) 2:5 [A-l] r-5r where IL? total length of pores with mean radius ?, ESP-0r = surface of pores covered with multi-layer thickness tr’ c = correction factor for the curved surface The volume of pores with radius ? is given by: -2 V? . U r 1‘? [A-Z] 167 so that L; = V1411 “f2 [A-3] Also the capillary surface 31' is given by: St = 2n rL? - 2%]? [A4] If equations (A-3) and (A-4) are substituted in equation (A-1) and the summation sign is replaced by an integral sign: 5"?)z m r-cr 2v? VYOr .- -—-2-—- VYOr + 0t 1‘ "f- T (It [A-5] ‘? ?+0r 'i’ - t... where c = ‘f t , and V?6r is the total volume of pores in the range or considered. In the limiting case, where 5r~0, equation (A-S) becomes: (r-tr)2 ”0° ”#:4er a A-6 vrdr 7— Vrdr + dt Jr r [ 1 where vr derived from experimental measurements, while r, tr, dr and dt are all functions of pressure which can be evaluated. Thus, Vr can in theory, be evaluated by applying equation (A-6) to the experimental results. In practice, however, it is not convenient to use the equation as it stands and it is preferable to integrate it over small finite ranges of radii. Consider a finite adsorption step from pressure P to pressure 1 P2, where P1 corresponds to the critical radius rl and P2 to radius r2. The total volume of water adsorbed during this step is: 1'2 v12 a I vrdr I1 2 r (r-t ) V (Zr-t -t ) 2 l " r l 2 _ Vrdr + (cz-cl) Jr 2 [A 71 l r 2 r «J~ 1' 168 E ,IHIIw'liltll....| .11 «A 11111 1.1.! II 1.. 11.111111171111941 {in}! 1...!!51111311333 .. . ”'iliififligl’hifliififiiijifiifilh’lflfliflfififES