-v\' f»), v n; . 5 --<-‘- ~‘ unev- ‘ vw.§v¢ 4,9" #.- ‘u . «11-- if an up u 1-" - .‘ ....- - no 9 '1' ”"9" Ir." 9.9M.“ ...“-.".II"I>' .t'. . .: ““3, W- "p .~ mu) g '?r H ,,.. .. ‘l _ o . . f. . .._, - J" 712?, h . ' L $4 ‘, ‘ .33. b") k V“ ” .~"‘r“"‘ ;.v'.-‘ » ‘ 23.3.3. Mm: 5:“ {1. £..-:.‘, , . :5, .. 3|". '3 u . :.'.- . ‘ “3.395.; 1‘": 1“ lg r U ‘ . ‘ 1_‘ '1. 1‘ J ' ' ' n. “1 7 J .r I) . 4&4! w." ‘ _ ..: 11“.! “4'11“ 9' *= in: s ‘ N .a . {-351.- figfggmggr. g; ’53}; :3 («$31 I ‘ 4,. r fi‘Pa’{~”“~‘ 'rr‘ #391”.§fi1,)’9§%§§h§§f§f% . .3: . f»: .x. - .‘3 *1 .\. ¢ :; 5‘; ; A. b' -a‘-‘t~ ‘ . , J) =.- 1”; ‘_ ,“A . 1 ‘ 1 }\“hp§1~q Q, f‘“ {:1}: l . t . E: .33}, ‘ ',-', h.” :Jj "" " ’ ”'"Z‘R‘W'U‘f‘éa "‘WE‘W‘} ‘ 1k! ’u t Weinaai .Ew! ;;a . - .- aka-b.3932“ . :g-tf’i}: ‘ Tfi'riiidrm 'E‘ Er _-.','i x 13-1! , - ,QI5fi5%$ . $2333.11, :4: Ufiqeitjs :34 ‘t' z‘ "i " ‘ ' J»? flak?“ ‘2' ‘F‘d‘ .: w “w z ' .m. . w - .A ‘<,I.-. .;; a” c "'"i'finqnhi ‘fi.:\ 1;}. lags . Ir ‘ .f‘}!::j“.{$"%‘ H “13.535 "x " 1: k‘ I .7 '. . Ab" ml: pfyrvtifyfjg" . 2 -: ~‘i= , --' psi-2m}; :. . a x.) '1 I. *9 ~ . mus .u.. 0N . U a. u: w I n 3| ""52“. ’ n.» '.‘ . i"? 0.; Kandy?! .. .. . _. . ‘r"3:,.-,~"'. . ,. 1'” ; (.51.? "’. t . , . Al"? ,3 ~..! ~ I _ Lo ‘7 ' ' ‘. ~"L :s ,-,. ~¢—~ 25' uncrpuu-u- A a» ‘. .‘ .... a . A- A, “H...“ h “A” - HD- Ou‘r‘flit 21¢?!- .. :7. .. 7'51: \mv". “- ~— - v; .m J a t-Aum1p' ‘ u.— «v _' n 1...:- w H - x M '_-. 43:3 “.4- »' nus... . 4. s -W——o.~—v Onc—v ’ "IV: V. . ‘331'93'5444. ,3. p“%“‘.] a}. ’. '1’ H m «m “4%; "j! 4 t «It y .- -A-q-o-. w. pug- .0- -\~ v... ' 33 . ”-0-; . .“,.,~.‘, ..—.~. ”5.5...- e a“; l 4:. - ’ ' a §hre32§fi ,E. '3", '11?! I 'f D ) .7“ - M'” .f}: “~53; H’ ' u .9..— I E Wt?!“ 91 " .. 3w 1‘; :4 :u-ah’uJXh‘i‘: a} ' -. a»... run..- mm 0 v “M m;- - A I.“ w: c— — 1' 2 L -= 1 "a. N’W Immmmmammm1mi'mTv’1i'TThImI 2’1 0 I? 786 3 1293 00602 2606 LIBRARY Michigan State University This is to certify that the dissertation entitled The Solubility of Nonpolar Gases in Organic Liquids and Water presented by Richard Paul Kennan has been accepted towards fulfillment of the requirements for Ph.D. degree in PhYSiCS 2 g Major professor Date March 9, 1990 MS U is an Affirmative Action/Equal Opportunity Institution 0-12771 4 ‘ A ~ « .64-.. -‘ PLACE ll RETURN BOX to roman this checkout from your toooni. TO AVOID FINES return on at baton one due. DATE DUE DATE DUE DATE DUE #6 MSU I: An Alfirmotivo ActloNEquol Opportunity Institmion The Solubility of Nonpolar Gases in Organic Liquids and Water. BY Richard Paul Kennan A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR 0F PHILOSPHY Department of Physics and Astronomy 1990 In @059 ABSTRACT SOLUBILITY OF NONPOLAR GASES IN ORGANIC LIQUIDS AND WATER. By Richard Paul Kennan We have measured the Ostwald solubility L(T), as a function of temperature in the approximate range 10.0-50.0°C, for 133Xe gas in 45 organic solvents, viz., 16 alkanes, 13 alkanals, 6 carboxylic acids, 4 alkanals, 3 cyclaalkanes, and 3 perfluaroalkanes. From our data for each salute-solvent system we determine the following . . . . . * thermodynamic functions of solution: chemical potential Ana: -* -¥ * -* -RTOnL, enthalpy Ah2 , and entropy Asa, where Auz= Ahz- -¥ TAsz, all based on the number density scale. The average observed entropy of salvation of Xe is A§:= -4.1 :t 0.5 cal/moi K, remarkably independent of solvent. The results are analyzed with scaled-particle theory from which we obtained the effective hard core diameters a‘, and the cavity energies gcnand enthalpies he" for all the solvents at 25°C. Thermodynamic perturbation theory is used to find the total enthalpy of salvation for the Xe-alkane systems. We generalize the analysis to evaluate the salvation enthalpy for all of the noble gases in the alkanes. We discuss the role of configuratianal entropy, as well as molecular dynamics approaches to calculation of free energies of salvation. Finally the results are examined empirically and values are given for the contribution to chemical potential, enthalpy, and entropy of salvation, of the six functional groups: CH2 (linear molecules), CH3, OH, COOH, CHO, and CH2(cyclamalecules). We have also measured the pressure dependence of the Ostwald (L) and male-fraction (x2) solubilities for the nonpolar gases N2, Ar, Kr, and Xe in water at 25.0°C. The pressure ranges studied for each gas were approximately: N2(44-116 atm), Ar(22-101 atm), Kr(33-81 atm), and Xe(5-48 atm). For N2, Ar, and Kr we see clear deviations from Henry’s Law, f2=knx2. The data are analyzed in terms of the Kirkwaod-Buff solution theory. The role of solvent -induced (hydrophobic) interactions shall be discussed” For the Kr-water system we shall compare our experimental results to recent computer simulation predictions. We also use statistical mechanics arguments to introduce a new solubility parameter which is appropriate to high pressure solubility measurements. Extensions of our analysis to other gas-liquid data is discussed. To Mom and Dad iv ACKNOWLEDGEMENTS The experiments described in this thesis were suggested by Professor Gerald L. Pollack. I cannot overemphasize his contribution to this work through his guidance, encouragement, and friendship. I would also like to thank Professors Jules Kovacs, Aaron Galonsky, Dan Stump, and Jerry Cowen for serving on my guidance committee. I would like to thank Jeff Himm and Gary Holm for their assistance in the laboratory in the early and late stages of this research project, and Charles Scripter for all of his help with computers. Many of the figures in this thesis were produced with the generous aid of John and Linda Kennan. Thanks are also due to Peggy Andrews for her general assistance. It would have been impossible to complete this work without the help of my friends both in the department and outside it. For moral support above and beyond the call of duty, I wish to express my deepest gratitude to Carrie Huang, Joel Gales, and Diandra Leslie-Pelecky. On the home front I would like to thank Richard and Nancy Olsen-Harbich for their kind friendship over the years. I also would like to thank the Physics Department's brilliant dramatic ensemble, the N.R.F.T.C. Players. Most of all I want to thank my family, especially my parents for whom I have seldom done my filial duty. The research reported here has been supported under the Navel Medical Research and Development Command. vi Table of Contents Chapter Page 1. General Introduction and Overview ................ 1 2. Introduction: Solubility of Xenon in 45 Organic Solvents ........................................ 3 3. Theoretical Background .......... . ................. 6 3.1 Phase Equilibria ....................... 6 3.2 Two Component Systems ...... . ........... 8 3.3 Ostwald Solubility ..................... 9 3.4 Salvation Thermodynamics .............. 10 3.5 Physical Interpretation...... ...... ...13 3.6 Other Concentration Scales ............ 17 4. Experimental.......... .......................... .19 4.1 Outline of Method ..................... 19 4.2 Radioactive Tracers ....... .. ......... .22 4.3 Solution Components ..... . ............. 23 4.4 The Experimental Apparatus............27 4.5 Volume Determination ................ ..28 4.6 Data Acquisition............ ..... .....33 4.7 Experimental Procedure ................ 36 5. Resu1ts..00000000000000.0.000...0.0.0.00000000000'040 5.1 Thermodynamic Analysis. ............... 40 5.2 Entropy of Salvation .................. 49 5.3 Enthalpy of Salvation.... ..... . ..... ..54 5.4 Temperature Dependence of L and Au....56 5.5 Total Entropy and Enthalpy............57 5.6 Male Fraction Scale ................... 63 6. Theoretical Analysis. ................ ...........65 6.1 The Excess Chemical Potential ..... ....65 6.2 Distribution Functions................70 6.3 The Van der Waals Picture of Liguids..80 * 6.4 Determination of Aug..................87 6.5 The Scaled Particle Theory ............ 93 6.6 Application to Real Liquids..........101 6.7 Evaluation of the Cavity Terms.......106 6.8 The Interaction Term.................134 6.9 Comparison with Experimental Data....146 7. Empirical Analysis and Predictions.. ..... .......159 8. Conclusions.. ................................... 171 vii 9. Introduction: Solubility of Inert Cases in Water at Elevated Pressure ....... . ........... .178 10. Experimental ................................... 181 11. Results ........................................ 188 12. Pressure and Concentration Dependence .......... 194 13. Data Analysis .................................. 201 14. Number Density Scale ........................... 211 15. Conclusions .................................... 219 Appendix A. Derivation of Equation (6.78) ................... 222 8. Computer Program to Evaluate ULM ........ . ....... 225 List of References .............................. 229 viii List of Figures Figure 1. Schematic of two component solution ............. 11 Figure 2. Three phase system of xenon gas in equilibrium with octane and water at 20 OC ................. 11 Figure 3. Schematic description of the salvation process. First the center of mass of the solute is placed at fixed position R0, and then the particle is released. The contributions to the chemical potential are indicated next to each arrow ................ 16 Figure 4. Idealized solubility apparatus.... .......... ....21 Figure 5. Experimental apparatus.......... ............ ....31 Figure 6. Schematic of detection electronics ...... . ....... 32 Figure 7. Normalized counting rate vs time for a typical experiment (Figure 2 of reference 18). The letter A indicates the time the valve was opened, and the letter B indicates the time a run might end, about 8 hours after equilibrium has been reached ..... 39 Figure 8. L(T) for Xe in six representative organic solvents ............................................ - ...... 48 t Figure 9. Excess chemical potential, A“; (T) for six representative organic solvents (Figure 1 of reference 17).. ................. .. ......... . .............. 50 Figure 10. Radial distribution function, g(r), for a typical gas, liquid, and solid..... ................. 75 Figure 11. Lennard-Jones potential, U(r). Lower diagram shows repulisive portion of U(r) as determined in WCA theory ................ . ................. 82 Figure 12. Radial distribution functions for (i)Lennard-Jones fluid g(r), (ii) for repulsive portion of L—J potential g°(r), (iii) for hard spheres gh.(r) ....... .... ................................. 83 ix Figure 13. Definition of a cavity in the SPT .............. 94 Figure 14. gcav(R12)/kT Xi cavity radius R12 ............. 100 Figure 15 Effective hard sphere diameter for 45 organic solvents 1; number of carbon atoms ............ 111 Figure 16. (a) gcav for Xe in 45 organic solvents at 25.0°C. (b) corresponding excess chemical 1: potential, Au: = —RT£aL .................................. 115 Figure 17. Molecular diameter ii polarizability for several inert gas solutes. Solid line shows extrapolation for noble gases ............................ 116 Figure 18. Excess chemical potential 11 solute polarizability for noble gases in Tetradecane, Decane, and Hexane. Filled in points along ordinate are the SPT predicted values for a hard sphere solute of 2.58A .......................................... 117 Figure 19. Excess chemical potential is solute polarizability for noble gases in polar solvents methanol, and ethanol. Filled in points along ordinate are the SPT predicted values for a hard sphere solute of 2.58A .................................. 118 Figure 20. (a) hbav for Xe in 45 organic solvents 11 number of carbon atoms.(b) corresponding * experimental excess enthalpy, Ah2 ....................... 123 Figure 21. (a) scav for Xe in 45 organic solvents 1; number of carbon atoms. (b) corresponding * experimental excess entropy, Asz ......................... 124 Figure 22. Heat of vaporization for 45 organic solvents at 25° C ys_number of carbon atoms ............... 125 Figure 23. Number Density for 45 organic solvents at 25° C _1 number of carbon atoms. ....................... 126 Figure 24. Thegmal Expansivity for 45 organic solvents at 25 C is number of carbon atoms ............... 127 Figure 25. Cavity free energy, gc , g§,T for Xe 0V in 6 representative organic solvents ..................... 133 Figure 26. Lennard-Jones potential with hard sphere cutoff ............................................ 138 Figure 27. Verlet-Weiss and Percus Yevick pair distribution functions for hard sphere fluid. Packing fraction = 0.5............... .................... 139 Figure 28. Calculated and experimental excess enthalpies of salvation for Xe in alkanes ................ 147 Figure 29. Calculated and experimental excess enthalpies of salvation for He, Ne, Ar, Kr, and Xe in (a) octane, (b) tetradecane. ........... . ........... 156 t Figure 30. Excess chemical potential Apz=-RT¢nL 21 number of CH2 groups at 25°C ........... . ............. 160 Figure 31. Experimental data for Xe in 37 solvents. Excess chemical potential divided by solvent number density is number of CH2 groups at 25°C.. ... ............. 163 Figure 32. Experimental data for Xe in 37 solvents. Excess enthalpy divided by solvent number density y§_number of CH2 groups ......................... 164 Figure 33. Experimental data for Xe in 37 solvents. * TAs2 divided by solvent number density is number of CH2 groups ............ . ...... . ................. 165 Figure 34. giM/p‘ y; number of CH2 groups for Xe in 37 solvents......... ....... .... .................... 170 Figure 35. Schematic of experimental setup (a) high pressure equilibration system (b) analysis system..................... ...... ..............182 Figure 36. Solute fugacity y§,mole- -fraction solubility at 25°C.... ..... . ......... . .................. 192 Figure 37. Henry' 3 constant y§,mole- -fraction solubility at 25° C............................ .......... .193 xi Figure 38. &n(fz/xz) - Pvzo/RT y§_mole—fraction solubility at 25°C ....................................... 203 Figure 39. 6n(f2/x2) — Pvzo/RT y§_mole—fraction solubility for He and N2 at 25°C for various values of Va0 ............................................ 210 xii List of Tables Table I. Solubility data for experiments with “BXe in 45 organic solvents. First row gives Ostwald solubility and the second row is the * excess chemical potential, Apz= -RT¢nL (cal/mol) ........ Table II. Experimental data for excess enthalpy, * * Aha, and entropy A52: for 1“Xe in 45 organic solvents .................................... . ........... Table III Enthalpy and entropy of solution. The second and third columns are the total entropy and enthalpy of solution. The fourth and fifth columns are the entropy and enthalpy evaluated on the mole fraction scale .............................. Table IV. Heat of vaporization, number density and thermal expansivity for 45 organic solvents ......... Table V. SPT calculations for 45 organic solvents. The second column is the effective hard sphere radius. The third column is the packing fraction. The fourth column is cavity free energy for 1E'Xe ..... ... Table VI. Cavity enthalpy and entropy ................... Table VII. Parameters to evaluate interaction energy. Second column is hard sphere cutoff. Third column is the energy integral using the Verlet-Weiss distribution function. Fourth column is the resulting Lennard—Jones energy parameter for the pure organic solvent ...... ............. Table VIII. Predicted enthalpy of salvation for “Xe in alkanes. ......... .. Table IX. Excess chemical potential, enthalpy, and entropy for He, Ne, Ar and Kr in the alkanes, including hexane, octane, nonane, dodecane, and tetradecane ........................ ............ ......... xiii ..42 ..51 ..60 .108 .112 .128 Table X. g“ , h , s , and um. for He, V GOV CGV Ne, Ar, and Kr in the alkanes,inc1uding pentane through hexadecane ....................................... 152 Table XI. Empirical energy parameters for Xe solubility in 37 organic solvents ..................... 166 Table XII Predicted y: expfi£imental values of Ostwald solubility for Xe in selected organic solvents ......................................... 166 Table XIII. Experimental results for solubility of N2, Ar, Kr, and Xe in water at 25.0°C. The second column is the solute partial pressure. The third column is the solute fugacity. The fourth and fifth columns are the mole-fraction and Ostwald solubilities respectively. The sixth column is the new solubility parameter r. The seventh column is the excess chemical potential of the solute in the liquid. The eighth column is the experimental average value of partial molar volume....... ............. 191 Table XIV. Results of Kirkwood-Buff analysis. The third column gives the partial molar volume of the solute used in Eq.(12.l4). The fourth and fifth columns give the results for the KB integral Ga: and the log of ghe infinite dilution limit of Henry's constant, cha )....... ........................ . ........ 204 xiv 1 .Introduction The phase equilbrium problem is ubiquitous to modern (or any) life. Whether it is the delivery of oxygen from the lung to the blood stream, or an industrial extraction process to remove unwanted hydrocarbons from natural gas, or the removal of a nasty stain from your best sweater; we are . constantly confronted with the problem of how a substance is transferred from one medium to another. In general, when two phases are brought into contact there is an exchange of matter until the concentration in each each phase attains a constant value. To understand the physics that determines this equilibrium state is a difficult, but important task. The goals of the research outlined in this thesis are to understand the solubility of gases in liquids and to predict solubilities in general gas-liquid systems. In pursuit of this goal we have measured the solubility of simple gases in various organic solvents and water. The resulting body of data has been analyzed using current molecular theories based on thermodynamics and statistical mechanics. By studying gas solubility one may be able to build a theoretical and empirical framework that can serve as a basis for more complicated phase equilibrium systems. The thesis is divided into two parts. The first section 1 details experiments involving solubility measurements of the 133Xe in several homologous series of organic radioisotope solvents. The solvents are chosen to range from simple nonpolar liquids to more complicated hydrogen bonding liquids. The second section covers experiments on the pressure dependence of gas solubility in water. The solute gases are: N2 , Ar, Kr, and Xe. These gases are meant to represent a series of prototypical nonreactive solutes. The experimental results are applicable to problems of medical, industrial, and. environmental interest. Specific examples will be sited in the introduction to each section. Gas solubility can also serve as a 'probe in understanding intermolecular interactions. Due to the advances in computational techniques it has become possible 1 ’2 The resulting to accurately model pure liquids. thermodynamic properties obtained are sensitive to the intermolecular potentials chosen. Optimized potential functions are determined by varying parameters until the best results are obtained.3 By extending these ’computer experiments’ to dilute liquid mixtures it may be possible to understand the various factors which influence the potentials and place them on a more sound physical basis. Accurate experimental data (n1 well chosen solute-solvent systems would be useful towards this end. 2.1ntroduction: Solubility of Xenon in 45 Organic Solvents. This section of my thesis describes experiments which measure the Ostwald solubility (L) as a function of temperature of’ 133Xe in 45 organic solvents including; alkanes, alkanols, cycloalhanes, alkanals, carboxylic acids, and perfluoroalkanes. A great deal of progress has been made in understanding the physics of simple liquid-gas systems such as the solubility of an inert gas in its own liquid, or the solubility of one inert gas in the liquid phase of another 4 ’ 5 These experiments inert gas, e.g., Aer, KroXe, etc. are prototypical in the sense that they are the simplest gas-liquid systems that one can study. A logical next step is to look at simple solutes in more complicated solvents '(the case of simple solvents with complex solutes isn’t experimentally accessible: Just try dissolving hexane in liquid nitrogen). To this end we have chosen the inert gas xenon in the several homologous series mentioned above. Some of the reasons why Xe was chosen as a solute are: All the inert gas elements are monatomic and do not «basically interact with the solvents under conditions of ' 3 these experiments. Much is known about interactions and properties of these elements. Xenon has a commercially 33 . . 1 Xe, whose concentrations in the available radioisotope, gas phase can be easily measured. Liquids are often considered in two groups: water6 (and aqueous solutions) and other liquids (mainly organic). We have selected 45 solvents so as to bridge these groups. Most of our solvents are nonpolar molecules, some are medium chain length molecules with a polar head group, and CHBOH and HCOOH are small polar' molecules for' which hydrogen bonding is important. The obvious motivation for looking at homologous series is that one can try to spot trends that can be generalized to other systems. Practical aspects of this work stem from biological and industrial applications» IXenon. has several applications which are dependent on its solubility and diffusion. Because Xe is highly soluble in fats and relatively 1'33Xe is widely insoluble in aqueous solutions, the isotope used in nuclear’ medicine to study cerebral blood flow, pulmonary function etc. A further application of Xe is as an inhalational anesthetic,7 a property associated with Xe solubility in lipids of cell membranes.8 Since the mechanism of general anesthesia is not understood some workers in the field believe that Xe is the prototype anesthetic to study.9 Also, solubility properties of Xe under pressure are useful for studying decompression sickness and inert gas narcosis, two problems of deep sea diving.10 Finally, there are environmental and safety 1”Xe and other 11,12 questions associated with the emission of radioactive inert gases from nuclear reactors. There is presently a great deal of interest in developing synthetic blood substitutes (now euphemistically referred to as oxygen carriers in the industry). Perfluorocarbons and related compounds form the basis of many' new blood substitute candidates because they carry oxygen efficiently and do not induce an immune system response.13 Thus understanding gas solubility in simple perfluorocarbons and their analogous hydrocarbons would be quite useful. While a first principles understanding is beyond the scope of this work, we hope to develop empirical and analytic techniques that will allow the prediction of Xe solubility from knowledge of bulk properties of the solute and solvent. We aim ultimately to generalize, at least qualitatively, the solubility parameters obtained for Xe to solubility of other inert gases. 3 .Theoretical Background 3.1 Phase Equilibria Chemical potential (u) is a fundamental quantity in the determination of phase equilibria in multicomponent systems The most useful definitions for our purposes are:14 a G ' a A u‘ 3 —— I . 3 (301) a ntlnr,n' a'ni'ny,n' where G is the Gibb’s free energy of the system, A is the Helmholtz free energy. n‘ is the number of molecules of type i,‘T is the absolute temperature, P is the pressure, V is the volume and N’ represents all other molecules in the i‘h type. system with the exception of the Suppose we have a two phase system, for example a liquid in contact with its own vapor, the condition for equilibrium at constant T and P is: ”('(TJ’) = u9(T.P) - (3.2) We shall use the subscript L to denote the liquid phase and q to denote the gas phase. For low pressures we assume the gas phase is ideal, ignoring internal degrees of freedom we have: p9 = -kT m (pg A3) = -kT On (p9 A3 / kT) , (3.3) where p9 is the number density of the gas (molecules/unit volume), It is Boltzmann’s constant and A is called the thermal wavelength for a particle of mass m; i.e. A=h/(2nka)1/z- The chemical potential for the liquid can likewise be evaluated: 3 * where pt denotes the number density of the liquid and Ap*is called the excess chemical potential. It is the contribution to the free energy from intermolecular interactions in the liquid. A more detailed derivation of the above equations is reserved until chapter 6, but for now we shall content ourselves with these results. Equating. (3.3) and (3.4) we find: * = -kT s 305 Au On (p, / pg) ( ) Thus by knowing the vapor pressure and density of the liquid we can determine its excess chemical potential. Typically, far from the triple point,a liquid is about 1000 times as dense as its vapor. This would lead to an excess chemical potential of about ~17 kJ/mole at room temperature. Although 8 this is a reflection of the strong binding energy in the liquid we must keep in mind that there are also entropic contributions to the chemical potential. These can play a large role as we shall see later. 3.2 Two Component Systems The relations ¢derived in the previous section can easily be generalized to two component gas-liquid systems. We now imagine a liquid solvent in equilibrium with a gaseous solute (Fig. 1). The condition for equilibrium is that the chemical potential of the dissolved solute equals the chemical potential of the solute in the gas: u; = H? . (3.6) Throughout this thesis the subscript 1 and 2will refer to the solvent and solute respectively, and superscripts 9 and C will refer ,respectively, to the gas and liquid, according to the usual conventions. The statistical mechanics of solute -solvent mixtures 15,16 starts with a standard partition function from which one may obtain by standard techniques the following chemical potential for a single solute molecule in the liquid solvent14’l7: u; = -kT a. < exp(-B° / kT ) > + kT on p; A3 .(3.7) 9 The key quantity in Eq.(3.7), Bo, is the binding energy of a single solute molecule to a fixed configuration of the solute-solvent system. The ensemble average of the exponential in Eq.(3.7) is taken over all such configurations. This term is also referred to as the excess chemical potential for the dissolved solute, All;- The second term in (3.7) is equivalent to the chemical potential of an ideal gas at density pf (the number density of solute molecules dissolved in the solvent). For the chemical potential of a solute molecule in the gas (assumed ideal) one has: pg = kT a. (93 A3) . (3.8) in which pg is the number density of the gaseous solute molecules. Equating (3.7) and (3.8) we find: * _ L 9 M2 - ‘kT 1CD ( pz / 92 ) o , (3.9) 3.3 Ostwald Solubility In the experiments described in this thesis we directly measured the Ostwald solubility as a function of temperature L(T) of xenon gas in various organic solvents. Ostwald solubility is an intuitive as well as a theoretically significant measure of solubility. It is defined as the (equilibrium) ratio of the (volume) concentration of dissolved gas molecules in the liquid solvent to their 10 concentration in the gas phase. If p; , pg are the number densities of solute 2 in the liquid and gas phases respectively, then: L = pg / pg . (3.10) At 20°C the Ostwald solubility of Xe is about 0.12 in water6 and about 4.4 in n-octane,18’19 a common nonpolar solvent (see Fig. 2). One can immediately see the importance of the Ostwald solubility by substituting Eq. (3.10) into (3.9), which gives: Au; = -kT Ln (L) . (3.11) The Ostwald solubility provides a direct measurement of the excess chemical potential. 3.4 Solvation Thermodynamics We start this section by reviewing standard thermodynamics. The Gibbs free energy for a system can be expressed in terms of its state variables aszo: G(T,P,N) = E + PV - TS = H - TS , (3.11s) l' G(T,P,N) = Znipi , (3.11b) m (9000 Figure 1. Schematic of two m‘9g95 C) 808%)! [Xe] =1 gas [Xe}=4 octane [Xe}=0.1 water 11 . solute o solvent component solution. [1 = concentration (dimensionless) T=20 ‘2: Figure 2. Three phase system of xenon gas in equilibrium with octane and water at 20°C. 12 where E is the energy of the system, S is the entropy, H is the enthalpy (H=E+PV), and n1 is the number of molecules of type i (of which there are r different species) whose chemical potential is pi. The differential is: r dG = -SdT + VdP + Z “idni . (3.12) -1 The most important information we shall be concerned with is the first derivative of the free energy, i.e: as s: -[———] , (3.13a) p we may then evaluate H in terms of G and its temperature derivative by using Eq.(3.11a). The partial molar enthalpy and entropy (h, and s respectively) can be found by differentiating (3.11b): [——-—a u‘] h T (3 13b) 8. = _ , o = u. + S. ’ O 1 a T 1 1 1 where ° r S = nsSI H = “oh. 0 (3.13C) .1 1 1 , .1 1 1 Generalizing these, we may find the partial molar enthalpy and entropy of the solvation process by taking the temperature> derivative' of the «excess chemical ‘potential, Aug: 13 6 Au *_- 2 *_ It * As2 - [ ) , Ah2 - Apz + TAs2 . (3.14) 6 T It * We call Ah2 and A32 , respectively, the excess partial molar enthalpy and entropy. One may also find the Helmholtz free energy of salvation and the internal energy of solvation. However these differ from G, and H by a factor of Pv2 (where vzis the molar volume of the solute in the solvent) which is usually negligible [see section 3.21 of ref. 14 for a complete discussion]. 3.5 Physical Interpretation of the Salvation Process In section 3.2 we introduced the concept of the excess chemical potential, ALI;- I shall now present a physical interpretation of this quantity developed by Ben-Naim.21 For simplicity we consider a two component system at temperature T, and pressure P with N1 and N2 representing the number of molecules of solvent and solute respectively, Because the Gibbs free energy is an extensive quantity the mathematical derivative in equation (3.1) can be replaced by: “2 = G(T,P,N1,N2+1) - G(T,P,N1,N2) . (3.15) This statement is valid for a macroscopic system, where the addition of one molecule may be viewed as an infinitesimal change in the variable ii In order to interpret various 2. contributions to the chemical potential Ben-Naim introduced 14 the concept of the pseudo-chemical potential. It is defined as: ”Escudo: G(T,P,N N +1; fio) - G(T,P,N1,N2) . (3.16) 1’2 This corresponds to the chemical potential for placing the solute atom at a fixed position within the solven. We assume the solvent is homogeneous and macroscopic so the actual position no is irrelevant as long as it is within the bulk of the solvent. By explicitly solving both (3.15) and (3.16) from partition functions one can show21: pg = "gum” + kT bu (pg/13) . (3.17) Therefore it is immediately apparent that the pseudo-chemical potential is equal to the excess chemical potential, A“; defined in Eq. (3.7). We may interpret equation (3.17) as follows. The full chemical potential can be viewed as a two step process for adding an extra particle to the solvent. First, we place the molecule at a fixed position. The change in free energy for this is “3"“d° (= Au; ). Next, we release the constraint imposed on the fixed position; this leads to an additional change in free energy, kam (pzAa). For classical systems 92A3<<1, thus the free energy resulting from the release of the particle is always negative. This quantity is referred to 21 as the liberation free energy. An explanation of the 15 various factors that make up the liberation free energy are as follows. When the particle is released it acquires a translational kinetic energy which leads to a free energy of kT bn(A3). Also, when the particle is released it may now wander throughout the volume of the solvent, which gives rise to a free energy -kT On(V). Finally, once the particle is released it is indistinguishable from the other N2 particles in the solvent. This adds a free energy contribution of kT (11(N2). Putting all of these together forms the liberation free energy. The entire discussion is outlined in Figure 3. The important property of equation (3.17) is that it is generalized for any kind of molecule, whether it be atomic argon or a complex protein, all that is required is that classical statistical mechanics be obeyed. Of course there still is the problem of developing a solid theoretical calculation of Au; , however we now at least have some feel for what it is we are dealing with. 3.6 Other Concentration Scales The Ostwald solubility, which has dominated much of our previous discussion, is based on the number density scale. This means that it is dependent on the number density of the solute in both the liquid and gas phase. However, there are many other measures of gas solubility. The most popular alternative is the mole fraction scale. The mole fraction solubility (xi) is defined as the eqilibrium ratio of the 16 ® ® @ kT in( p2 Ag) Figure 3. Schematic description of the salvation process. First the center of mass of the solute is placed at fixed position 30, and then the particle is released. The contributions to the chemical potential are indicated next to each arrow. 17 number of moles of dissolved solute divided by the total number of moles of solution: x. = ——————' . (3.18) For a two Component solution this reduces to: = N2 / (N1 + N (3.19) 2) 9 where N2 and N1 are the number of moles of solute and solvent respectively. Note that unlike the Ostwald solubility this makes no direct reference to the gas phase concentration. If the gas phase is ideal we can relate x2 to the Ostwald solubility, L19: -1 ._. 1[ R_T_ . 1] , (3.20) L where P2 is the pressure of the solute gas and 171 is the molar volume of the solvent. For a dilute solution the chemical potential for component i may be written as 22: “i = p: + kT &n(xz) . (3.21) The quantity u: in equation (3.21) is known as the standard state of the dissolved solute. It is hypothetical in the 18 sense that we cannot give a physical interprtation of it. For example, one might say that it is the chemical potential of the pure solute (i.e. x2=1), however equation (3.21) only holds in the case of very dilute solutions. Despite this flaw the mole fraction scale is still useful and has many 20’22 Its strength lies in the fact that it can applications. be used to describe liquid mixtures and electrolyte solutions, where knowledge of the vapor phase is often hard to obtain. There are of course many other concentration scales that 20 but only the number density scale avoids are available, the problem of defining standard states and allows a strict physical interpretation of the associated chemical potential. 4. Experimental 4.1 Outline of Method For the discussion in this section. please refer to Figure 4. Here we have a two compartment chamber separated by a valve. The upper chamber contains the solute gas while the lower chamber contains the liquid solvent and a-stirring device. An accurate pressure gauge is connected to the gas volume which will allow us to determine the gas density at any time through its equation of state. The gas volume in the upper chamber is denoted V“’, the liquid volume is Vt , and the gas space in the lower chamber is V‘z’ 8 system is immersed in a constant temperature bath. The . The entire (1) V' 8 amount of gas. This is determined by measuring the initial procedure of the experiment is to fill with a known pressure, determine the gas density from it, and multiplying the density by the initial volume: (1) Nuu41.1:"hnieun ' Vg ' (4'1) We then open the valve and start the stirrer. The pressure will drop until equilibrium is reached and the solvent is saturated. We now measure the final pressure and determine the final density, (For this idealized system we pfinel. 19 20 neglect the vapor pressure of the solvent . In a real experiment we must take the solvent vapor pressure into account.) 'The final number’ of solute particles can. be expressed as: _ gas (v(1)+ V(2)) + dissolved: V Nfinel- final)‘ 8 8 pfinel AC ’ (4.2) where the second density is that of the solute which is in the liquid. We now can introduce the Ostwald solubility, which for this system is obviously: dissolved final L = 9.. . (4.3) pflnel We now substitute L * p9" for p“"°”°d in equation final final (4.2). Since the system is closed the number of solute molecules is conserved, thus we can equate (4.2) with (4.1) and solve for L: (I) (‘1 (2) v v 4 v L = a -3— - -3——-3- , (4.4) "’4 V: where a is the ratio of initial to final gas density, (pi/pf). To briefly summarize, we have outlined a simple technique in which L is obtained by recording the decrease in pressure during equilibration. 21 Pressure Gauge valve (1) Figure 4. Idealized solubility apparatus. 22 4.2 Radioactive Tracers There are several problems which come up when actually measuring the Ostwald solubility as described in the previous section. First of all, one must be sure to measure the pressure drop associated with the solute gas. Two contributions of background gases are, (a) the vapor pressure of the solvent, and (b) any other gases dissolved in the liquid. For dilute solutions the vapor pressure of the solvent does not change significantly from its pure value,15 so it can be corrected for. The second problem is overcome by carefully degassing the solvent when loading it into the apparatus. This can lead to fairly elaborate and time consuming setup procedures. Another problem is the necessity of noninvasive pressure measurement. In other words, we do not want to change the system by measuring its pressure. Many of the more accurate pressure gauges and sensors are able to detect changes in pressure through moving parts such as diaphragms or Bourdon tubes.23 A result of this is often a small change in gas volume which must be accounted for due to our need to know absolute differences in mass (density*volume). The most difficult problem to overcome is sensitivity over a wide range of gas concentration. As mentioned earlier, gas solubility in various solvents can vary by more 6’24 The best reasonably priced than factors of one hundred. transducers and gauges have an uncertainty of 0.05 percent of the full scale reading. Therefore a 0.05% error at one 23 pressure could become a 5% error on another. To get around this one must use a series of pressure sensors for the range appropriate to the system being studied. A related difficulty is that for systems where the Ostwald solubility is low, the observed change in pressure will be small compared to the total pressure (or even the vapor pressure). This can be overcome by using large liquid volumes, however for some expensive solvents this is not a feasible alternative. By using radioactive tracers one can avoid all of these problems. Since the tracer is the only gas which is observed, background gases can be ignored. Furthermore, the tracer can be monitored by external detectors, which are truly noninvasive. And finally, since radioactive decay has an uncertainty of (NcmmflJ)-o.5 one can get desired accuracy over a wide range by simply waiting for enough counts. Since the goal of this experiment is to look at gas solubility over a wide range of solvents, the advantages of using radioactive tracers is apparent. 4.3 Solution Components The solute gas chosen for our experiments is the radioisotope Xenon-133. Xenon is one of the noble gases. Its atomic radius is 2.23A, which qualifies it as the largest and most polarizable of the 5 inert gases with stable isotopes, viz.,He, Ne, Ar, Kr, and Xe. The next noble gas in this series is radon which has no stable isotopes. 24 235U fission and is readily Xenon-133 is a byproduct of available commercially. It is unstable and decays with a half life of 5.245 days.25 Because of its high solubility in fats and low solubility in aqueous solutions, 133Xe is used extensively in nuclear medicine to study cerebral blood flow pulmonary function, etc.18 The decay process for’133Xe is as follows: the isotope first decays by beta emission to an excited state of 133C 3, this nuclear excited state then decays with a half life of 6.3 * 10"9 sec by emiting an 81 keV gamma ray. The beta rays are rapidly attenuated, but the gamma ray intensity can readilybe quantitatively 133 measured to determine xe concentration. Xenon-133 was purchased from the Hedi-Physics Company (Plainfield, N.J.) in 20 millicurie (mCi) aliquots (this is a recommended human dosage for cerebral blood flow studies). A typical amount of ‘33 Xe used during a run was of the order of 100 pCi. In practice, one aliquot usually supplied enough xenon for a month (a: 10 runs). The corresponding partial pressure of the tracer is appoximately' 1‘ picoatmosphere (10"12 atm), therefore» it’s safe to say that our results correspond to the limit of infinite dilution. The tracer was usually mixed with air at 1 atm. In order to make sure the air did not effect our results control experiments were done in which' 1”Xe was mixed with naturally occuring nonradioactive xenon at a total pressure of 1 atm. No difference was observed. The solvents studied were an extension of previous work involving n-alkanes and 25 n-alkanols26 (the prefix n- means that the molecule is a straight chain). They include cycloalkanes, alkanals, carboxylic acids, and perfluoroalkanes. N-alkanes can be described as straight chains of singly bonded carbon atoms which are saturated with hydrogen: H H H C H 9 H-C-C-eeeoe-C-H e (405) n 2n+2 H H H They are composed of n22 acyl groups (CH2) and two methyl groups (CH3). We studied the alkanes ranging from pentane (C5) to eicosane (020). The alkanes were purchased from Humphrey Chemical Co. (New Haven, Conn.), and were all at least 99% pure. N-alkanols differ from n-alkanes in the addition of a hydroxyl (OH) group to a terminal carbon: , H "' eeeee " C - 0H s (406) H“ - C nH2n+l 1 2 H H OH 9 H - C - C H H Ethanol was obtained from Aaper Alcohol and Chemical Co, while the other alkanols were from Aldrich Chemical (Milwaukee, Wis.). Purities were: 99.9% (methanol), 99+% (propanol and butanol), 99% (pentanol, octanol, decanol, and undecanol), 98% (hexanol, heptanol, and dodecanol), 97% (nonanol and tetradecanol), and 200 proof ethanol. Carboxylic acids are formed by replacing the terminal hydrogen with a carboxyl group (COOH): H H H on H2n+lcmn ’ H - 1.01,- :2- eeeee "" g“- COOH e (4e?) Alkanals, also known as aldehydes, have a terminal CHO group (CnH2n+1CHO)' Cycloalkanes (CnHzn) are similiar to alkanes except that the carbon atoms form a closed ring; as a consequence they have no methyl groups. These solvents were also obtained from Aldrich Chemical. Their purities were the highest reasonably available: formic acid (95%-99%, remainder water), acetic acid (>99%), propanoic acid (>99%), n-butanoic acid acid (>99%), n-pentanoic acid (>99%), n-heptanoic acid (>99%), propanal (>99%), n-butanal (>99%), n-pentanal (99%) , n-heptanal (95%), cyclopentane (78%; 99.6% saturated 05 hydocarbons), cyclohexane (>99%), and cyclooctane (>99%). Perfluoroalkanes are alkanes with fluorine substituted for hydrogen (CnF2n+2)° They were obtained from SCM Specialty Chemicals (Gainesville, FL). Their purities were also the highest reasonably available: perfluorohexane (99%, of which 85% is n-CGFIO), mixed isomers), and perfluorooctane(90% mixed isomers). perfluoroheptane (97%-99%, Other perfluoroalkanes were either prohibitively expensive or not available. 4.4 The Experimental Apparatus A diagram of the apparatus is shown in Figure 5. The design is similiar to that used in previous work,26 except 27 the upper and lower portions are now held together by an indium sealed brass flange instead of a threaded brass connection. The apparatus can be disassembled at the flange for cleaning and loading. The upper portion of the apparatus consists of two valves and a brass gas volume. The lower portion is a pyrex flask which is joined to a glass-to-metal seal. The brass flange is soldered to this seal. A small matching groove is cut on both faces of the flange and is filled with 0.040" diameter indium wire. The top and bottom are connected by six screws which press the indium to form‘s reliable seal. The indium can be reused many times by reforming the wire in a hydraulic press. The lower volume contains a glass encased stir bar in order to mix the solvent. Most commercial stir bars are coated with Teflon because it is non-reactive. We found it necessary to remove this coating because xenon is very soluble in Teflon. The stir bar is then encased in glass to protect it from potentially corrosive solvents. A ball valve separates the upper and lower volume. The advantage to using a ball valve is that its volume is well defined. In other words, once the valve is opened the volume that is exposed is that of the hole in the ball regardless of how much you open it. The second valve, a Hoke valve, is for loading the gas. This does not have any critical volume requirements because it remains sealed throughout the entire run. During a run the apparatus is immersed in a fluid bath 28 in order to regulate temperature. The bath was controlled to $0.1”: by a Lauda/Brinkman K-2/RD circulator. The bath fluid was a 4:1 mixture of water:ethylene glycol in order to prevent freeze-up of the circulator cooling coils. We used Pb sheilding a 1 cm to reduce background radiation and to isolate the volume Vé1)from the rest of the apparatus. This is important because we want the detector to measure the concentration in the gas phase only. Any scattered gamma emissions from 133Xe in the liquid which reach the detector will lead to errors. One cm of Pb 6 27 attenuates BlkeV gamma rays by a factor of 10 We checked the shielding by putting ‘33 Xe in V with the rest apparatus in the bath and the shielding in place. No counts were observed above background. 4.5 Volume Determination One of the° most important criteria for determing solubility by the method outlined in section 4.1 is the accurate determination of volumes.‘ The relevant volumes for the apparatus are indicated in Figure 5. Of these seven volumes five are the same for every run, namely, V‘1’,V g rest (2) , V;b, Vfib , and V.. . The remaining two , V8 and Vt , are determined at the start of each new run. Each volume is described below: (1) V 8 = the rest of the gas volume , = initial gas volume , (2) V 8 29 = total volume excluding V(1) , rest 8 Vt = volume of the liquid , Vhb = volume of hole in ball valve , V“ = volume of stir bar , V = volume below the ball valve . They are related by: = V + V (408) rest ee hb (2) ’ v8 — rest. vsb VC . (4.9) To determine V" (ee stands for everything else), we weigh the apparatus before and after filling V“ with water at a known temperature. We make a buoyancy correction for the weight of air displaced by the water. If mob. is the observed mass, then the true mass,muu.,is: p sir,T mum: mob. / [1 - Fla—7;] , (4.10) 2 where p 'r is the density of x at temperature T. One can X, then determine V" by: Vee= mtrue / pl! 0,1' . (4.11) The volume of the hole in the ball was calculated from measurements made on the disassembled valve. We can then 30 find V}..t from (4.8). The volume of the glass encased stir bar, V.b, was found by measuring the water it displaced in a graduated cylinder. Once V"... was known , we were able to find v; 1 ) by performing a dilution run. This consists of loading 1331Xe in Véi) with V}..t evacuated. Once we determine the concentration of gas in Vé" , c°(normalized to correct for radioactive decay), we open the main valve and allow the gas to expand into Véz’. After several hours the system equilibrates and we note the final normalized concentration, ct. Since the amount of xenon is conserved (subject to radioactive decay) we have: (i) (1) cov8 - f(V8 + Vr.‘t) , (4.12) which in turn gives: V“’ = V s a / ( 1- a ) , '(4.13) 8 rest 31 ; Nal a I Detector ] I I 5 5 . . I I Pb Shielding ’ g \ I 1 I I I- a I I 5 5 I I I I Valve a v1' V1 .- g ' Pb Shielding / Valve H L—I ‘ ' brass flange glass-metal seal "—’ 2 V ,_. 9 l Viiq ' l stir5ar Figure 5. Experimental apparatus. 32 High Volltage Power Supply 0 d Signal Single etector an . —. a Amplifier Channel Photomultiplier P" Amp fl Pre- Amp Analyzer Power sand . AOOSec ._.—4 _ . Timer _ Counter Tngger Connections 4 Counting 200 Sec __ EM” ‘ ' Com uter “m" Timing Pulse p CRT Printer Figure 6. Schematic of detection electronics. 33 where a = (C0 / of). For the apparatus used in these measurements Vuus 250cm3 , and Véus 50cm3. 4.6 Data Acquisition The electronics are shown in Figure (6). Gamma ray intensity is monitored by a well -counter NaI(T1) scintillation crystal connected to a photomultiplier (Harshaw Chemical Company, type 7SF8). A high voltage power supply (Power Designs, Inc., model 1570) supplies 1100 volts for the photomultiplier via the preamp (Ortec model 276). Signals from the photomultiplier go through the base preamplifier and a Nimbin mounted amplifier (Ortec Model 485), then through a single channel analyzer (SCA, Ortec Model 406A), and finally into a microcomputer counter (Micro Development Tool 1000). The SCA is set with a suitably wide window (10-90keV) so as to permit passage of ‘33 Xe signals while keeping the system stable and the background low. The window was set by feeding signals of known amplitudes into the amplifier and observing the output on a multichannel analyzer (Northern Electronics model NS633). The resulting 133Xe spectrum . Two t imers output was then compared to a (Ortec Model 719) are supplied a 0.1 sec timing pulse by the computer. The computer is programmed to take data and print out the results. It counts the number of decays in a 400 sec interval and then waits an additional 200 sec before counting again (the time intervals are somewhat arbitrary 34 and are based on historical reasons). A result is printed every ten minutes with averages printed every hour. Two corrections are made to the raw data. Because the electronics are not perfect, a pulse into the detector is not recorded if it follows an earlier pulse too closely, i.e, the detector has a ’dead time’ The dead time can be measured by sending signals of known frequency into the electronics and observing the output. In this way, we determined the dead time for our system to be about t a: 3.33xlo—6 sec. In order to calculate the dead time correction we must exploit the random nature of the radioactive decay process. The probability of k events in a time interval, given a mean rate of u events in that interval, is described by a Poisson distributionza: k u p(k) = . (4.14) kleu We now take the dead time, t, to be our unit of time. Thus p becomes the mean number of events in time t. The probability that a single signal occurs within time t of an earlier signal is given by: 9(1) = u e'" . (4.15) Since p is small (typically of order 10-3) we can expand the exponential. To first order this reduces to: 35 P(l) u . (4.16) The actual counting rate is then approximated by: Nact = Nobs(1 + P(1)) = Nob.“ + u) , (4.17) where Now. is the observed number of counts over period T. We approximate u by using the observed counting rate multiplied by the dead time: u = ( Nob./ T ) * t . (4.18) Plugging into (4.17) we have: Naet = Nob.( 1 + ( Nob./ T ) * t ) . (4.19) We have programmed the computer to make this correction using the parameters : T=400sec, and r=3.33x10-63ec. The second correction we make is to subtract off background counts. We measure background radiation over periods of several hours with the evacuated apparatus and shielding in place. Background is subtracted after the dead time correction is made. Typical background rates are about 1500 counts/400 sec. During an actual run counting rates vary from 250,000 counts/400 sec (at the beginning) to 50,000 counts/400 sec (at the end). 36 4.7 Experimental Procedure I shall now outline the experimental procedure to 33 . . 1 Xe in an arbitrary determine the Ostwald solubility of solvent. Referring to Figure (5) may be useful. The solvents are generally used as purchased, we do not degas them. Control experiments were done on degassed solvents and have shown the dissolved gases do not effect 26 xenon solubility ( if we equilibrated the solvent with 1 atm of air, the gas concentration would amount to about 0.01 mole/liter). With the ball valve closed and the apparatus (1) v 8 loading valve (Hoke bellows valve). A trace amount of disassembled, air is evacuated from volume through the 1’33Xe is allowed to expand into V”’, and then air is let in 8 to bring the total pressure to 1 atm. The loading valve is then closed. The glass portion of the apparatus is loaded with about 200 ml of solvent, and is weighed (we make the usual buoyancy correction).' Using known densitieszg’30 we calculate the liquid volume. The perfluoroalkanes and cyclopentane were mixed isomers, so we measured the density ourselves using a standard volumetric method. This allows us (2) 8 the stirring bar, the apparatus is assembled (see Section to calculate gas volume V using (4.9). After putting in 4.5). The apparatus is then mounted rigidly in the temperature bath so the detector geometry does not change in 37 the course of the run. With the main valve closed, we start the computer and begin taking data. During a run the initial counting rate, ci, is measured hourly for six to eight hours. Each value is corrected for decay, and the mean value is calculated. (1) g 9 by a steady counting rate, we open the main valve and start Once the gas is uniformly distributed in V as indicated the stirrer. The solubility can then be calculated from a generalization of equation (4.4): V(1)+ V(2) Vr where 1 is the decay constant for 133Xe (l=0.9177x10-4min-1) and At is the time elapsed between measurement of c1 and cf. We calculate the Ostwald solubility each hour until L reaches a constant value. Time to equilibrium is typically 12 hours. We wait an additional 8 to 12 hours to insure equilibrium has been reached(see Figure 7). In order to obtain L(T) in these experiments one needs to know the densities o(T) at each temperature at which L(T) is measured. The idea is that after charging the apparatus with solute and solvent and then measuring L at some initial temperature, say 25°C, one can measure L at a different temperature by just changing the temperature and waiting for the new solute-solvent equilibrium. Once the bath temperature is changed we recalculate equation (4.20) using 38 the new liquid and gas volumes, VC and V’Z). This procedure 8 works because the masses of solute and solvent are fixed in the sealed apparatus. Thus, in a typical determination of L(T) we measured solubility sequentially at temperatures of 20,30,40,50,10 and, finally, again 20°C (the temperatures we cycle through depends on the solvent of course). We went though a cycle like this two times independently for each solvent. 39 5C) . I [if I I {rfrr I i [i I rl i I r [r‘ a—40u '- g A. _ c . 3 . >20 . - b O E . " a: . h, . _ in —- .3 . _ 7 '. ‘- se' - - .< 5 " a)4 -, .. '3 B X3. l- 3... i o ’ f 1‘ 6 3 ' ‘4 Time (hours) Figure 7. Normalized counting rate vs time for a typical experiment (Figure 2 of reference 18). The letter A indicates the time the valve was opened, and the letter B indicates the time a run might end, about 8 hours after equilibrium has been reached. 5. Results 5.1 Thermodynamic Analysis Table I gives the results of our experiments for 45 solvents. These include: 16 alkanes, 13 alkanols, 6 carboxylic acids, 4 alkanals, 3 cycloalkanes, and 3 perfluoroalkanes. The first row for each solvent gives the measured value of L(T) at each temperature. The second row gives the corresponding chemical potential, Au; (cal/mole), which is calculated from L(T) by equation (3.11): Au;=-RTan . The data given at each temperature are each averages of at least 2 separate runs. Typical uncertainties for 6L/L are about 0.015. The uncertainty 6(Au;) = -RT(6L/L) varies from £6 cal/mole at 278.15K to :10 cal/mole at 323.15K. Figure 8 shows the Ostwald solubility plotted as a function of temperature (C) for six representative solvents, one for each homologous series. They are: n-hexane, n-hexanol, cyclohexane, n-heptanoic acid, n-heptanal, and n-perfluorohexane. Each of these solvents has a similiar backbone with different terminal groups attached (or hydrogens replaced by fluorines). Figure 9 is a plot of the 40 41 corresponding chemical potential versus temperature. For the solvents illustrated the largest solubility occurs in hexane, this corresponds to the most negative free energy, while the lowest solubility occurs in perfluorohexane, which has the highest free energy. From these figures we can see the general tendency in the Xewmu- organicwlvmt system for L(T) to decrease with temperature, while ApZCT) increases with temperature. From equation (3.11) it is clear that the sign of Au; is negative if L is greater than one, as is the case for 44 of these solvents (HCOOH excluded). For all such solvents a positive free energy is required to remove a solute molecule from a fixed position in the solvent to a fixed position in the gas. Conversely for those in which L is less than one, a positive free energy is required to place a solute at fixed position in the liquid. Over the temperature range we used, the experimental data for Ap;(T) can be fitted well by a straight line of the form. Ap;(T)=a+bT. Since the partial molar entropy is proportional to the temperature derivative of the chemical * potential, 3: = -(a"2/6T)p , one can “has.“ -RT on (L) = Au;(T) = All: - ME: , (5.1) 42 gable I. Solubility data for experiments with Xe in 45 organic solvents. the Ostwald solubility, L(T), First row gives and the second a): row gives the excess chemical potential A“: = ~RT¢nL. Temperature is in degrees celsius. Temperature 10 20 30 40 50 alkanes n-carbon 5 L(T) I 6.41 5.48 Aule) I -1045 -991 6 L(T) I 5.91 5.07 4.55 Au:(T) I -999 -945 -913 7 L(T) I 5.41 4.67 4.13 3.75 3.37 t AMI(T) I -950 -896 -854 -822 -780 8 L(T) I 4.99 '~4.36 3.90 3.47 3.31 a Au.(T) I -904 -856 ~820 -774 -769 9 L(T) I 4.70 4.14 3.70 3.32 2.99 s Aule) I -a71 -828 -788 -747 -703 10 L(T) I 4.42 3.92 3.52 3.14 2.84 t Auzfl‘) I -836 -796 -758 -712 -670 ll L(T) I 4.16 3.72 3.35 3.00 2.71 8 Aung) I -805 -765 ~728 -684 -640 12 L(T) I 4.03 3.59 3.22 2.90 2.64 a AM:(T) I -784 -744 -704 -662 -623 13 L(T) I 3.88 3.44 3.09 2.80 2.53 I Au.(T) I ~763 -720 -679 -641 -596 14 L(T) I 3.76 3.35 3.02 2.72 2.49 a Au.(T) I -745 -704 -666 -623 -586 15 L(T) I 3.24 2.92 2.64 2.41 c Au'lT) I -685 -645 -604 ~565 43 Table cont..... 16 L(T) 3.14 2.35 2.57 2.35 Aule) -667 -631 -537 -549 17 L(T) 2.75 2.51 2.30 Au:(T) -612 -573 -53s 13 L(T) 2.71 2.47 2.25 Au:(T) -601 -563 -521 19 L(T) 2.42 2.21 . Au'(T) -550 -509 20 L(T) 2.36 2.17 . ““2‘7’ -534 -493 Table I cont..... 44 Temperature 10 20 30~ 40 50 alkanols n-carbon 1 L(T) - 2.46 2.20 1.93 1.79 Au:(T) . -507.4 -460.4 -411.5 -362.3 2 L(T) - 2.79 2.47 2.22 ' 2.02 1.35 Au;(r) . -577.9 -527.9 -431.5 -436.0 -395.4 3 L(T) - 3.02 2.65 2.33 2.16 1.93 Au:(T) . -621.3 -567.7 -521.3 -432.2 -439.0 4 L(T) - 3.04 2.63 2.40 2.17 1.93 * AnalT) - -625.3 -574.7 -527.4 -432.2 -439.0 '5 L(T) - 2.97 2.62 2.36 2.13 1.95 3 Aule) . -613.3 -561.3 -516.5 -467.0 -419.6 6 3(3) . 2.97 2.62 2.34 2.12 '1.92 . . Auz(T) . -611.9 -559.1 -512.7 -467.0 -419.6 7 3(7) - 2.91 2.57 2.31 2.09 1.91 Ap;(T) - -601.6 -594.4 -503.3 -457.5 -413.5 3 L(T) . 2.36 2.52 2.25 2.05 1.33 Ap;(T) a -590.7 -537.3 -439.6 -445.3 -404.4 9 L(T) - 2.79 2.49 2.24 Au;(T) a -577.5 -530.5 -434.5 10 L(T) . 2.74 2.43 2.20 2.00 1.33 AuZlT) . -567.3 -513.4 -475.5 -432.0 -333.4 11 L(T) . 2.34 2.11 1.92 1.76 Au:lT) - -496.2 -449.3 -406.3 -363.3 12 3(7) - 2.12 1.94 1.73 Au:(r) - -453.5 -410.3 -369.6 14 L(T) - 1.91 1.72 Au;(T) - -404.0 -364.3 Table I cont..... 45 Temperature 10. 20 30 40 50 carboxylic acids n-carbon 1 3(7) = 0.474 0.456 0.437 0.420 0.414 Au;(T) - 420.1 453.1 493.7 539.3 566.3 2 3(7) . 1.71 1.53 1.47 1.37 Au;(7) - -312.0 -275.2 -239.7 ~204.3 3 3(7) - 2.97 2.66 2.41 2.21 2.04 i 332(7) - -611.7 -570.0 -530.0 -493.2 -456.6 4 3(7) - 3.22 2.39 2.61 2.37 2.16 0 Aule) a -653.7 -613.9 -577.0 -537.5 -494.3 5 ' 3(7) - 3.20 2.34 2.56 2.31 2.11 I Aule) . -653.9 -607.4 -567.2 -521.3 -473.6 7 3(7) . 3.12 2.73 2.50 2.27 2.07 R “"2‘T’ - -639.5 -594.9 -553.0 -511.0 -463.1 alkanals . 3 3(7) 2 3.09 2.30 2.55 fl 332(7) - -634.9 -600.0 -564.4, 4 3(7) - 3.47 3.12 2.33 2.56 2.40 * 332(7) - ~700.0 -662.5 -625.6 -535.7 -560.9 5 3(7) - 3.50 3.15 2.34 2.59 2.40 Au;lT) - -705.2 -663.3 -623.4 -592.4 -560.0 7 3(7) . 3.32 2.93 2.70 2.47 2.26 I 332(7) . -675.7 -636.3 -593.4 -561.9 -524.2 46 Table 1. cont.... Temperature 5 10 15 20 25 n-carbon perfluoroalkanes 6 L(T) I 2.48 - 2.39 2.30 2.20 2.11 Aule) I -501.8 -489.8 -477.9 ~459.3 -442.4 7 L(T) I 2.23 2.16 2.09 . 2.01 1.95 Aule) I -443.3 -433.8 ~423.2 -406.7 -396.3 8 L(T) I 2.15 2.06 2.00 1.92 1.85 * Aule) I —422.1 -408.0 -396.3 -379.7 -365.8 Temperature 30 40 50 n-carbon perfluoroalkane 7 L(T) I .l.87 1.76 1.67 e Ap2(T) I -378.0 -350.4 -327.8 8 L(T) I 1.79 1.67 1.59 a AM2(T) I -349.4 -318.0 -296.6 47 Table I cont..... Temperature 5 10 15 20 25 n-carbon cycloalkanes 5 L(T) I 6.88 6.49 6.04 5.75 5.33 Ap:(T) - -1066 -1053 -1030 -1019 -991.3 6 L(T) I 5.28 5.00 4.70 Aule) I -952.2 -937.3 -916.8 8 L(T) I 4.26 4.02 3.79 Au:(7i . -329.4 -309.9 -739.7 Temperature 30 40 50 n-carbon cycloakanes 5 3(7) - 5.04 Ap;(T) a -974.3 6 3(7) . 4.46 3.93 3.53 ‘R 032(7) = -9oo.1 -359.7 -313.5 3 3(7) - 3.53 3.21 2.91 Ap;(T) . -763.3‘ -725.3 -636.2 Ostwald Solubility, L 48 I I I I @ T 1 I I FT T l I I l I I I [ I I fit : O Alkanes : _ X Alkanols o 9 3 To Acids ‘j o - 0 Cyclo 8 I : >< Perfluoro : —+ Aldehydes o -— I— -l - 0 _. + ..l .— D -4 ._.- )1 + —-( .— . D .4 _ X n a 1 4. — X -( _ X X x D J» L— X n .ifl x l— .7 .4 l l l L l ' l l J I i L L l I L i L l I l i l i O 10 20 30 4O 50 Temperature, °C Figure 8. L(T) for Xe in six representative organic solvents. 49 i.e., the slope of Au;(T) is the negative average entropy of solvation (AEZ) over the temperature interval, and the intercept at T=0 (T is the absolute temperature) is the average enthalpy of solvation (Ahz). One might at first think that this interpretation ignores the effect of a strong temperature dependence of the enthalpy having a strong temperature dependence which could then be seen in a: Au2(T), however this is not correct since31 (6H/6T)p" : T(aS/6T)p" . (5.2) Any temperature dependent contribution from the enthalpy to the free energy is cancelled by a temperature dependent contribution from the entropy (provided we take the derivatives under the proper conditions). This does not mean that aAh:/8T is meaningless, in fact this quantity is 14 This quantity turns the specific heat of solvation, Ac:. out to be very small, but it may be of interest in more detailed studies in the future. The tabulated values of Ah; and 3;: for the 45 solvents studied can be found in Table II. 5.2 Entropy of Solvation In viewing Figure 9 one can see the slopes of the straight line fits are all quite similiar. This means that the solvation entropies all fall close together. This Chemical Potential. Apao' (cal/moi) 50 -4£"3 r'Ii l ’ Izii" ii’i i‘ i ii I ‘i r*ii l r l l l « - 3“ 500 b «c 3“ 0% _. -— ,0 ‘° _ ‘9 (jfififi -600 —- o _, _ «63,903 : —700 -- "j -300 —— - —900 *- -1000 -— _fi11c"3L- 1 1 LJ lllllLII Ill [I I (I LUlJ_L- 0 10 20 30 40 50 Temperature. C t Figure 9. Excess chemical potential, A“2(T) for six representative organic solvents (Figure 1 of reference 17). 51 Table II. Exggrimental data for excess enthalpy and entropy for Xe in 45 organic solvents. alkanes n-C Afi:(;:t) A§:(%§%x S ~2582 -5.429 6 -2264 -4.482 7 -2109 -4.121 8 -2109 -4.260 9 -2044 -4.l45 10 -2007 -4.128 11 ~1958 -4.069 12 -1922 ~4.017 13 -1924 -4.104 14 -1879 -4.005 15 -1869 -4.038 16 ~1819 -3.928 17 -1769 -3.820 18 ~1807 ”’ ”.3393— 19' -l762 ~3.875 20 -1727 -3.808 L. -. cycloalkanes n-C 36333;) A§;(§§{K‘ 5 -2120 -3.77 6 -2065 -3.85 3 -2022 -4.14' 52 Table II. cont ..... alkanols n-C Ah:(§§%) A§:(%§a‘ 1 -1330 -4.34 2 -1369 -4.57 3 -1910 -4.57 4 -1943 -4.66 5 -1924 -4.64 6 -1959 -4.77 7 —1922 -4.67 3 -1900 -4.64 9 ‘ -1390 -4.65 10 -1326 -4.45 11 ~1790 -4.41 12 -1720 —4.20 14 -1630 -3.90 Table II. cont.... 53 carboxylic acids 46:42:) :<—- 1 -640.2 -3.75 ’-2 -1366 -3.60 3 -1702 -3.86 __; -1319 -4.10 5 -1335 -4.35 7 .134, -4.26 alkanals (aldehydes) n-c Alibi-2t) 3%.: 3 -1632 -3.52 4 ~1710 -3.57 5 -1743 -3.67 7 -l742 -3.77 perfluoroalkanes n-c 9532-3?) 653.53%; 6 —1335 -2.99 7 ~1188 -2.67 8 -1221 -2.87 54 observation is further supported on Table II which indicates -* that the values of Asz are all negative and commensurate in magnitude. The average value of the salvation entropy over all our solvents is, (A§:)‘v. -4.1t0.5 cal/mol K . The entropy is associated with the process of taking the solute from fixed position in the gas and placing it at fixed position in the solvent. The negative entropy can be thought of as solvent ordering during this process. We can conclude from our data that the amount of ordering is more or less independent of the solvent used. This isn’t too surprising since our solvents were generally large floppy molecules compared to the solute. We shall see later that the salvation entropy is much more dependent on the solute particle. It is interesting to compare our results to Xe solubility in water. For the Xe-oHZO system, the corresponding A52, averaged from 10-40°C is about -18 cal/mol K , calculated from the data of Benson and Xrause.32 This is indicative of a high degree of solvent ordering upon hydration. This ordering can be associated with the high polarizability of Xe, hydrogen bonding, dipole-dipole interactions, and the small relative size of H20 molecules.9 These «combined effects can lead to ice-like structures 9,32 surrounding the solute , and thus decreases the entropy of the liquid. 5.3 Enthalpy (Energy) of Salvation Before analyzing the enthalpy a brief reminder, for the 55 systems we are looking at the excess enthalpy of salvation is essentially the same as the excess energy of salvation since PV is small in a liquid. If we assume the gas phase is ideal then the salvation enthalpy, Ahz, can be thought of as the energy required to take the solute from infinity and put it at a fixed position in the liquid. One can see from Figure 9 and Table II that for all the solvents, except HCOOH (formic acid), the salvation process is enthalpically dominated, i.e. AfileA§:> 1. For these solvents (excluding HCOOH) we find AhZ/TA§:=1.510.15, also remarkably constant. The average value of enthalpy for the same solvents is A52: -1840 :1: 250 cal/mole. Values for enthalpy are spread more widely than for entropy as one can see from Figure 9. The alkanes and cycloalkanes tended to have a large negative enthalpy, while the perfluoroalkanes and the small polar molecules (such as acetic and formic acids) have enthalpies of (comparatively’ small. magnitude. All values of enthalpy were negative. This demonstrates a net attraction between the solute and the solvent. Solubility of Xe and other noninteracting gases in water is entropically dominated. Using the XeIHZO data from Benson and Krauseaz we find that 36:: -4050 cal/mole, averaged from 10-40°c, which leads to Afi:/TA§: = 0.75. Even though the enthalpic contribution is larger in water than in the organics it is overwhelmed by the entropy caused by solvent organization. It shouldn’t really be surprising that the enthalpy in water is high since water has a permanent 56 dipole) moment, but the interactions between. other water molecules are also very strong. This competition is reflected in the salvation entropy and keeps the solubility of nonpolar gases in water low. A simple way to think of it is that water likes the solute, but it likes itself a lot better and rearganizes to keep as many hydrogen bonds intact as possible. 5.4 Temperature Dependence of L(T) and Ap;(T) It is clear from our earlier discussions that the temperature dependence of the chemical potential is determined by the entropy. Thus we attribute the positive slope of the curves of Figure 9 to a negative entropy. We now ask: What determines the temperature dependence of the solubility itself, L(T). To illustrate this clearly we start with the now familiar relation (equation 5.1): * -* -* Au2(T) = -RT LniL) = Ah - TAs2 . Solving for L(T) we have: - 35* 35* 3(7) = exp [ 2 + 2 ) . (5.3) 37 R Assuming A32 and A3: are essentially constant in the temperature range of interest we differentiate (5.3) with repect to temperature: 57 -* 3n LL: 1,—22 . (5.4) dT R T 33 One can also show: -* -Ah2 (T-To) ) L(T) = L(To) exp [ . (5.5) R T T a where T is the temperature of interest and To is a chosen reference temperature. This leads us to the surprising ' result (at least for me it was) that the temperature dependence of the Ostwald solubility is due to the enthalpy, while the temperature dependence of the chemical potential is due to the entropy. We conclude that the experimentally observed decrease in gas solubility with with increasing temperature (see Figure 8) is associated with negative enthalpies of salvation. 5.5 Total Entropy and Enthalpy Up to this point we have focused on the entropy and enthalpy associated with the excess chemical potential, 50;. This is the chemical potential for the fictional process of placing the solute at fixed postion in the solvent. Its value in constructing a physical model of salvation has been discussed earlier. We would now like to look at the total entropy and enthalpy for the entire (real) salvation process. Our starting point is the chemical potential for 58 the solute and the equlibrium condition (equations 3.6-3.8): p9 = kT On (pg A3) 2 u: = kT On (p; A3) + An; 9 _ L “2"“2 ‘ Taking the temperature derivative at constant pressure and composition we calculate the total entropy: tot_ _ _§_ 4 _ 9 _ 3 _ 9 As2 - 8T [(12 112 ); [s2 s2) , (5.6) where: 8p“ 39 = -k on (p9 33) - kT -1— ——2 (5.7) 2 2 pg a'r P 4 6p 3: = -k Ln (9% A3) - kT[ —l— -—E ] +33: - (5.3) P For fixed particle number we can write the derivatives of density as negative derivatives of volume. This allows us to substitute the isobaric expansivity, a = (1/V)(aV/8T)P into (5.7) and (5.8). For dilute solutions we assume the expansivity of the liquid mixture is the same as that of the pure liquid, at . The gas expansivity, 019 , is easily derived from the ideal gas equation of state: 59 a l / T . (5.9) Putting all this together, the relationship between the a total entropy, As;°t , and the salvation entropy, As2 is 33‘“ = 33* - k can.) - k + kTa 2 2 C . (5.10) We can now find the total enthalpy through equations (2.6) and (2.13b): Aug“ = 0 = 311;“ - 733;“ . (5.11) Therefore: 31);“ = TAs: - 1:73:1(3) - kT + szoi£ , (5.12) which is equivalent to: 2 tot _ * _ Ah.2 - Ah2 + kT kT at . (5.13) tot ’°’and As2 Table III shows Ah2 for the 45 solvents used. Table III. 60 Enthalpy and entropy of solution. The second and third columns are the total entropy and enthalpy of solution. and fifth columns are the entropy and enthalpy The fourth evaluated on the mole-fraction scale. alkanes ze— e3— 5 —2899 -9.72 —2818 -16.90 6 -2615 -8.77 -2691 -16.37 7 ~2478 -8.31 -2415 -15.36 8 -2493 -8.36 -2429 ~15.34 9 -2440 -8.18 —2378 -15.09 10 -2410 —8.08 -2351 -14.93 11 -2379 —7.98 -2308 -14.74 12 —2348 -7.88 -2278 -14.56 13 -2350 -7.88 -2278 -14.51 14 -2307 -7.74 —2239 -14.30 15 ~2301 —7.72 —2233 -14.23 16 -2253 -7.76 -2188 -l4.00 17 -2206 -7.40 -2154 -13.68 18 -2246 —7.54 -2191 -13.91 ‘cycloalkanes 9:633 93:23.3 A5323) 3%..) 5 -2478 ~8.31 —2385 -15.68 6 —2437 ~8.17 -2263 -15.34 8 -2440 -8.18 ~24l7 -1S.84 61 Table III cont... alkanols “72%) A‘s';<:.:t.n A6:<:::> Aé:<:.::.( 1 -2263 -7.59 -2250 -18.82 2 —2268 -7.61 —2257 —17.87 3 -2325 -7.80 -2300 -17.39 4 ~2372 -7.96 -2345 ~17.l3 5 -2361 -7.92 -2331 -l6.79 6 -2397 —8.04 -2364 -l6.64 7 -2359 -7.91 -2317 -16.26 8 -2342 -7.86 -2304 -16.05 9 -2343 -7.86 -2310 -15.90 10 -2272 -7.62 -2234 -15.50 11 -2239 -7.51 -2210 -15.35 carboxylic acids n-C Arlyfi) A§;(;—:{-K) Ah:(;::) ":(Egt—K 1 -1052 -3.53 —1051 -17.99 2 -1780 -S.97 —l780 -l7.02 3 -2101 -7.05 -2088 -16.67 4 ~2223 -7.46 -2204 -16.49 5 -2301 —7.72 -2287 -16.48 7 -2276 -7.63 -2239 -15.84 62 Table III cont... alkanals (aldehydes) -t cal. -!. col. -0 cat -0 col. n-C Ahz(;3T) A52(mouc) Ah2(;3T)‘Asz(moLK) 3 -1966 -6.60 -l946 -l6.13 4 -2071 -6.95 -2039 —15.83 5 -2115 -7.09 -2076 -15.61 7 -2152 -7.22 “2137 -15.38 perfluoroalkanes -t. cal. -t cal. -0 col -0 col. n-C Ah2(;:T) 2(moLK) Ah2(;3T)‘Asz(moLK) 6 -l632 -5.48 -1606 -l3.5 7 -1515 -5.08 -l487 -l3.0 8 -1565 -5.25 -1536 -13.1 63 35,36(we Expansivity data was obtained from standard sources measured the expansivity far the perfluoroalkanes and cyclopentane using our temperature dependent density data, p(T) ). Although the salvation enthalpy, Ah: is not accessible to direct measurement, the total salvation enthaply, Ahgu, is. This is the heat of solution at constant pressure. The heat of solution has been measured for various gases in water using calarimetric techniques, and general agreement is found with solubility derived 34 values. No such measurements have been made to date on our system. 5.6 Male Fraction Scale The standard approach to solubility is to calculate chemical potential on the mole fraction scale Au; (see discussion in section 3.6) : o-— The corresponding enthalpy (A52) and entropy (A3:) are obtained by writing An; = Ah: - TAEZ and proceeding in a manner analogous to what we have used. The tabulated enthalpies and entropies can be found in Table III. Note that the enthalpy an the sale fraction scale (A52) is very close to the total enthalpy of salvation (Ah;°t). This fact has strengthened confidence in the mole fraction scale, but also note the entr0py (AE2) is not the same as the total 64 -* °‘) or the entropy of salvation (Asz). In fact entropy (As; the mole fraction entropy has no straightforward interpretation (and,in my opinion, is often misused). 6. Theoretical Analysis 6.1 The Excess Chemical Potential, Au;. Now that we have experimentally determined the excess chemical potential, also referred to as the free energy of salvation, our task is to predict it from as close to first principles as possible. This goal will constitute the bulk of this chapter. The chemical potential is derived from the partition function. In order to keep things clear we will first work out the case for a single component, N particle, atomic liquid in the canonical ensemble (all results are generalizable to the grand .canonical ensemble). The classical partition function is: Q(V.T) a —L—” I 6i" I d3" apt-BIG". i5")! . (6.1) N! h where l’is the hamiltonian of the system, ha' (h is Planck’s constant) accounts for quantum corrections to the differential volume elements in phase space, N! corrects for the indistinguishability of the particles, 8 is the Boltzmann factor (um, and (if-",3” is a paint in the an 66 dimensional phase space that describes the system.16 An important simplification that arises in the classical appoximation is the separation of potential and kinetic terms to characterize the liquid state.35 This allows us to write the Hamiltonian as: 2(3'.B" = K(B”) + U(B") . (6.2) where K(p') is the kinetic energy of the classical degrees I of freedom, K(B')=£ (pf/2m), and 0(3') is the potential is! energy. We can now write the partition function as: - 1 an an __ an Q(V.T) - [m h" Jdr Idp expl B K(p )1] x -l; I d?" expl-B U(f")] . (6.3) V . In writing equation 6.3 we have used the fact that [drul V" = 1 (note that we now have two integrals over d3"). The first bracketed term 'is the partition function for an N particle ideal gas, Qua-1° The second term is referred to as the configurational partition function , Qcon . 16 Evaluation of the ideal partition function is straightforward:36 A-au v” , (6.4) ideal- N1 67 A denotes the de Broglie wavelength, A=(2uBh2/m)”2. The Helmholtz free energy, A(N,V,T), is: A = -kT {n.Q(V,T) . (6.5) For our partition function this is equivalent to: A=-kT{mQ -kTOnQ =A +AA , (6.6) ides co 1 n ideal is the free energy for an ideal gas and AA: is I the excess free energy. The corresponding chemical where A ideal potentials are found by differentiating with respect to N, it therefore becomes obvious that the chemical potential can be expressed as: +Ap , (6.7) where u” = -kT 01(pzA3). To derive an expression for the eel excess chemical potential we use identity (3.15): Au*= AA*(N+1,v,T) - AA*(N,v,T) . (6.8) v econmwm.) ) = kT 6n [ Q (N+1.V.T) can The ratio Qc"(N+1,V,T)/Qcon(N,V,T) is given by: 68 Iexpl-flul.1(;"1)l d;u+1 = q . . (6.9) Qc°n(N.V.T) Iexpt-BUlh'H dr" Q (N+1,V,T) can If the potential energy is pair decomposable we may write the potential energy of the N+1 particle system as: +N+1 _ 9N U".1(r ) - Uu(r ) + a0 , (6.10) where mo is the binding energy of the (N+1)'t particle with all the others in configuration 3'. Equation (6.9) becomes: 1 exvl-Bool expl-BU'(?')J d?'*‘ = . (6.11) Qc°n(N.V.T) jexpl-BU'(3')1 a?” Q (N+1.V.T) can In a system with translational invariance, the point 1"” may be taken as the origin 'for the remaining N position vectors. 'This allows us to integrate over’ 3"‘, ‘which yields a factor of V. Equation (6.11) reduces to: Q nmfl’v’” v I ”pl-3%] eXPI-BU.(I-")l di’J' Qc°n(N,V,T) Iexp[-BU.(r')] d?" = V , (6.12) where the angular brackets denote a canonical ensemble average over the N particle system. Substitution of (6.12) 69 in (6.8) gives: Au*= -kT 6n . (6.13) For a two component system ‘we can write the partition function for a simple structureless solute (2) dissolved in a fluid (1) Q = Qid..l Qiczleal v-(N‘HNZ) x Id?"‘j6§'zexp[-p( 01(3"‘) + uigf",?"2) + uzgf'zil. (6.14) For dilute solutions the contribution from solute-solute interactions, 022 is negligible.14 Following the same steps as before one can show that the free energy for adding the (Nz+1)th particle is: Au; = - kT 6n , (6.15) where B0 is the binding energy of the added salute to a fixed configuration of the solute-solvent system, i.e: N291 +N2+i - N2 4N2 012 (r- ) — 012(r ) + Bo , (6.16) and the ensemble average is taken over all coordinates of the (N1+N2) particles. Now we can appreciate an important simplification that occurs by using an inert structurelesss solute like Xe as a 70 probe. If we used a more complicated molecule with internal degrees of freedom (such as vibrational and rotational) it is conceivable that the potential energy would be a function of the solute conformation (shape), we would have to perform a double ensemble average: one over all spatial configurations of the (N1+N2) molecules and a second over all possible configurations of the solute molecule itself, i.e: An; = -kT on < > (6.17) P i 2 P2 represents the internal coordinates of the solute L4 A clear example where this would be necessary molecule. is the salvation of a polymer in water. By using an inert gas probe we can focus on free energy changes due to the liquid solvent and not have to worry about contributions from the solute. 6.2 Distribution Functions An alternative approach to describing the liquid state that has proven to be quite powerful is the distribution function method. I shall present a brief overview of the technique, a more complete discussion is given in reference 36. As in the last section I shall work in the limit of an atomic liquid to keep the development simple,, but the results are readily generalizable. The phase space probability distribution function in a 71 classical system is defined as: expl-Bz(?'.§')l/ Idf'jdfi'eXPt-Bl(3'.3'). (6.18) The probabilty of a state (3",3') is simply f(?",3")d§'d§". We have shown earlier that the Hamiltonian factors into kinetic and potential terms. Therefore we can write the total phase space distribution function as a momentum probability distribution, M3"). and a configurational probability distribution, P(r'). The latter will be important in this discussion. The configurational probability distribution tells the probability for observing the system at configuration space point '1'", it is defined as: “-11-, = expt-BUG'H / I dr'expl-BUH’J'H . (6.19) Note the denominator is equal to (V'x Qcon). We) can. determine distribution functions for' a small number' of ‘particles by integrating over (all coordinates except those pertaining to the particles of interest. The joint probability distribution for. finding a labelled particle 1 at 31 and a labelled particle 2 at 32 is: (ZIN) 4 4 _ a 9 e an F (r1,r2) .. I dra I dr‘....I dr. P(r ) . (6.20) Since we cannot label identical particles, a more meaningful 72 quantity is the reduced probability distribution function, (2")(ifi 1 p ,fz). It is defined as the probabilty distribution for finding any pair of particles at fiand $2: -) _ _ (21):)» .. 1,rz) - N(N 1) I’ (r‘,r2) . (6.21) The factor of N(N-1) refers to the fact that there are N ways of choosing the first particle and (N-l) ways of choosing the second. Similiarly, the n-particle distribution function is defined as: I d?"'" expl-BU(§')] p(n/N)(;n)= N! (N-n)! . (6.22) I d?” expl-BU(?')] For an homogeneous fluid, the single particle distribution function is simply the bulk density: p(1")(;1)= p = N/ v . (6.23) In the special case of an ideal gas, U(f'")=0 and QC”. 1. Thus the n-particle distribution function becomes: N! p‘"""('§n) = p" m '= p"(1 + O(n/N)) . (6.24) For example the pair distribution function for an ideal gas is: 73 3 ,3 ) = pz(l - l/N ) a p2 , (6.25) (ZIN)( 1 2 p where the last equality depends on N being large; which it is for any macroscopic system. We now introduce yet another .9 distribution function , g(f1,r2): g(f .? ) = p p . (6.26) which is the fractional deviation of the two particle distribution from the ideal gas limit. It is essentially a measure of how much the system deviates from complete randomness.35 If the system is isotropic as well as homogeneous, the pair distribution function “3132) is a function only of the separation between the two particles, r12: Ifi- 32| ; it is then referred to as the radial distribution function and simply written as g(r).16 From our definition of g(r) one may write the following:' ps(r) = p(p(z’"’(0.i") / p 2) = p ‘2’") p . (6.27) Since we have already seen 9”“)(31) = p, equation (6.27) can be interpreted as follows; pg(r) is equivalent to the conditional probability density that a particle will be found at 1" given that another is at the origin. This is based on the well known theorem of statistics: If A and B are random 'variables *with. joint probability' distribution 74 P(AnB), then the conditional probability distribution, P(AIB), that A occurs if B also occurs 1837: P(A|B) = P(AnB)/ P(B) , (6.28) where P(B) is the probability distribution for B. Our definition of pg(r) can also be stated as follows; pg(r) gives the average density of particles a distance r from a particle at the origin. The average number of particles in volume element dV at distance r is thus: dN(r) = pg(r)dV . (6.29) Figure 10 shows g(r) for an atomic solid, liquid and gas. The radial distribution function plays an important role in the physics of liquids. There are two main reasons for this. First, the radial distribution function is directly measurable by scattering experiments (although it can be difficult to interpret results for complex molecular 75 0 1 L L 0 20 30 Fun . . cootdination she" Second COOl’dlllallOIl Sh!" The radial distribution function for a simplefluid. ‘Y l 4).. '*-—-Som1 3i— ‘ Lhum \\ 2” r) I ————— —— —— \ — \ 1 (I l l 1),, 0 la 20 Radial distribution functions for liqAUid and solid argon at the triple point (oi-34 ). Figure 10. Radial distribution function, g(r), for a typical gas, liquid, and solid. 76 liquids). Secondly, if the particles interact through pairwise additive forces, thermodynamic properties of the fluid can be written in terms of integrals over g(r). The two thermodynamic functions we shall require in our analysis are the excess internal energy of salvation, unit, and the excess free energy of salvation, Ad‘. I shall now present brief explanations of how these quantities are determined from g(r). Detailed derivations can be found in reference 36. The internal energy of an atomic fluid can be written I!) u NH» Cl NkT + (6.30) where the first term is the mean kinetic energy (same as for an ideal gas), and the second term is the mean potential energy. Using our definition of pg(r) we can easily find 6. Consider a tagged particle at the origin. If we assume the intermolecular interactions, u(r), are pairwise additive we can express the potential energy between the tagged particle and all other particles at distances between r and r+dr as: d0 = u(r)dN(r) = u(r)pg(r)dV = u(r)pg(r)4sr2dr (6.31) The total potential energy of the liquid is found by integrating over r, and multiplying by N/Z. The factor of N arises since any particle could be the tagged particle, and 77 the factor of two compensates for double counting of pair interactions. The average potential energy per particle is therefore: Cl ll 2|... m = 211p! u(r)g(r)r2dr . (6.31) o For the case of a dilute solution (we use the conventional subscript notation in which 1 refers to the solvent and 2 refers to the salute), the excess internal energy (excluding kinetic ener8Y) of the solute is: z duinr' u12(r12)p1g12(r12)4ur12 driz ’ (6'32) In Eq. (6.32) the function g12(r12) is the solvent-solute pair correlation function and p3g12(r) is the number density of solvent molecules a distance r from the tagged solute. The excess internal energy for this solute molecule is found by integrating over r12 0. uint- 4np1Io g12(r)u12(r)r2dr . (6.33) This is sometimes called the excess internal energy of salvation. Note that for a pure liquid, for which ughz} goes into “hid“, the excess salvation energy is twice the average excess energy per particle , 3. This is because the average energy in equation (6.31) is the potential energy to assemble. the whole liquid, while the salvation energy in 78 equation (6.33) is the potential energy to bring one extra particle into a pre-assembled system. In order to calculate the excess free energy of salvation we must introduce the idea, due to Onsager38 and Kirkwood,39 of a coupling parameter 5 which can vary from {=0 to 5:1. We imagine the following process: we would like to introduce a new particle into the system by turning on its interaction with the other particles in the system. When §=0 the new particle does not interact with the system at all, and when §=1 we have the full intermolecular interaction u12(r12). This can be expressed as: u12(§,r12) = Eutzhtz) . (6.34) To find the excess chemical potential we use equation (6.8): Ap* = AA*(N+1,V,T) - AA*(N,V,T) . Since the Helmholtz free energy is related to work in thermodynamics ‘we can conclude that the excess chemical potential is the isothermal, reversible work that has to be done on the system against intermolecular forces in order to add one more molecule to the system under conditions of constant volume and temperature.40 In other words Au* is the work done on the system in going from the initial state with N molecules coupled with each other and 1 molecule not 79 coupled (5:0) to a final state with N+1 molecules fully coupled (5:1). Following the same steps as before we choose the new molecule to be at the origin. For an arbitrary intermediate value of 5 we can write the radial distribution function about the central molecule as g(r,§). The potential energy of interaction of the central molecule is §u(r). This means that u(r)d§ is the work that is done on the system by this one interaction if E is increased by d5. The work done on the system when 5 increases by d5, due to all the molecules between r and r+dr is: du(r,§) = u(r)d§ pg(r,§) 4nr2dr (6.33) The total work, Ap*, is the integral of (6.33) over r and over 5 from 5:0 to 5:1: * 1 a 2 Au = 4np I I u(r)g(r,§)r drd§ . (6.34) a a For a two component system the salvation free energy is: _ * i a 2 Apz = 4uptj I u12(r)g12(r,§)r drdE . (6.35) o o A generalized form of this statement is: 1 Au; = (New: . (6.36) 0 80 where U(E) is the spartially averaged coupling energy of the salute to the rest of the system with the interaction potential reduced from its full strength by a factor of 5.41 6.3 The Van der Waals Picture of Liquids One of the fundamental problems in developing a theory of liquids is that there is no idealized model comparable to the ideal gas or the harmonic solid; both of which can be treated exactly.31’42 These provide a reference system from which one can base perturbation expansions or at least get an intuitive feel for the physical situation, in order to develop more sophisticated theories. The van der Waals picture of liquids has helped a great deal to overcome this problem. The basic idea is to look at the different roles of strong short-ranged repulsive intermolecular forces and longer-ranged attractive forces in determining the structure and dynamics of a dense fluid. Though the concept was first utilized by van der Waals in his treatment of nonideal gases, there have been many other contributors to our current understanding.43 The renaissance of this idea was spurred by the discovery from computer simulations that a system of infinitely hard spheres (essentially billiard balls, albeit tiny ones) undergoes a first order fluid-solid phase transition that can be related to freezing and melting of real materials.44’45 The van der Waals picture asserts that the relative arrangements and motions of molecules in a liquid are 81 determined ‘primarily by packing effects produced by the short-ranged repulsive interactions. Attractive forces, since they vary relatively slowly, play a minor role in structure and to first order can be treated as a mean field which exerts no intermolecular force but provides the cohesive energy that holds the system together at fixed density and temperature. As an example we shall briefly look at a Lennard-Jones fluid, which is defined by the intermolecular potential, u(r) (see Figure 11): u(r) = 4.:[w/r)12 - (a/r)6] , (6.37) the collision diameter, a, is the separation of two particles where u(r)=0, and e is the depth of the potential well at the minimum in u(r). The properties of, the Lennard-Jones .fluid are well known from computer simulations,46 and it serves as an excellent model of atomic liquids like argon.47 The repulsive (l/r)12 term arises from electron cloud overlap and Pauli exclusion, and the attractive (1/r)6 term accounts for induced dipole - induced dipole interactions. To examine the role of repulsive and Potential [ U(r) ] Potential [ U(r)] 82 0.501 U“) '8 0.0 1 .0 2.0 3.0 Intermolecular Distance (r) Figure 11. Lennard-Jones potential, U(r). Lover diagram shows repulisive portion of U(r) as determined in WCA theory. L00 4 0.50 3 1 A q V O 000 4 ~050‘ 4.00 ‘ ( -L50‘ 5U(r) --------)-------. ,4 C -200 fi- r I 00 L0 20 30 Intermolecular Dlstance (r) i g(r) 83 l k golf) | —-- g}\9(r) 3““ I. 000 g(r) 2»- ' . O 1- l- '.-— .--- -. .. -- l . l (3 __ "' d - 2d r Figure 12. Radial distribution functions for (ilLennard-Jones fluid g(r), (ii) for repulsive portion of L-J potential g°(r), (iii) for hard spheres gh.(r). 84 attractive interactions in determining liquid structure tests were done to compare g(r), the radial distribution function for the full potential u(r), to go(r), the radial distribution function that one would obtain from only the replusive forces (at the same temperature and density). The potential used to determine go(r) is (Figure 11): uo(r) = u(r) + e rsro u°(r) = 0 r>ro . (6.38) In Eq.(6.38) ro is the point at which the minimum of u(r) 21/6 occurs; i.e. r = a. Figure 12 shows the strong a correlation between g(r) and g°(r) as determined from computer simulations and analytic theory.48 To take this experiment one step further it was found that there exists a hard sphere system for which the radial distribution function, gh.(r), is closely related to the actual distribution function, g(r) (Figure 12).49 The hard sphere fluid is characterized by the sphere diameter, d, which appears in the hard sphere potential (Figure 11): ii I a rsd u = 0 r>d . (6.39) The effective hard sphere diameter is chosen to most closely reproduce the structural features of the real fluid. The fact that the attractive forces play a minor role in 85 structure was rationalized by Verlet.50 In order to understand his argument a few new points must be introduced. The local particle density is defined as: M?) = 2M '1’- - ii) , (6.40) where 8(x-xi) is the Kronecker delta function. The average density at point f is simply , which for a homogeneous system is the bulk density, p. The Fourier transform of (6.40) is: pk: I exp(-ik-f)p(f)df = S: exp(-k-f'i) , (6-41) is! with an autocorrelation function defined as: I sat) = T‘ pup > . (6.42) The function S(k) is called the static structure factor. For a homogeneous fluid it can be related to the Fourier transform of g(f) 35: SHE) = 1 + p I exp(-ik-'f') 3mg): . (6.43) S(k) can be experimentally determined from scattering 86 experiments on the liquid. One can then perform an inverse Fourier tranform to determine g(f').16 One can show that the limit of S(k) for k40 is 35: - _ 0 3(0) - kaxT - xT/ xT . (6.44) where xr is the isothermal compressibility of the liquid: - 1 23 IT " p [arr] 9 (6.45) N r and x: is the isothermal compressibility for an ideal gas: x2 = 1/ka . (6.46) If we apply an external field to the fluid with potential (M3), one can show that this leads to a Fourier transformed density response, 69* 35: 6pk = -S(R)¢(§)p/k7 , (6.47) where (HE) is the Fourier transform of Ni"). Equation (6.47) is a form of the more general fluctuation-dissipation theorem. Verlet’s arguement' can now be summarized. Equation (6.47) shows that the structure factor determines the density response of a system due to a weak, external field. If the external potential is associated with the potential 87 of a test particle at the origin, the long-range part of that potential gives rise to a long wavelength response in density. In the long wavelength limit (k90), the response is proportional to 8(0) by equation (6.47). By equation (6.44) we see that 8(0) is proportional to the compressibility» ‘Eor typical liquids, the compressibility is very small ( xT/ x: m 0.02 near the triple point). Therefore, the density change caused by a long wavelength perturbation is not significant. This phenomenon is referred to as ’repulsive-force screening’.51 At lower densities, such as in the critical region, the compressibility (and hence 8(0)) becomes large . ‘This can lead to large density fluctuations. In this regime attract ive forces become very important and the van der Waals model is no longer valid. 6.4 Determination of Au; , ’Hard’ and ’Saft’ Contributions The success of the van der Waals model in predicting structure has lead to the development of thermodynamic perturbation theories. The idea is to split the excess chemical potential into two parts: a reference term due to the ’hard’ repulsive interactions, and a perturbation term due to the ’soft’ attractive interactions: + An . (6.48) :7: M2 = A“ soft hard To treat the ’soft’ part of the potential we perform energy 88 averages using the structural information obtained from the ’hard’ part of the potential (this is analogous to quantum mechanical perturbation theory where we compute the expectation value of the energy by employing the wave function of the unperturbed hamiltonian). In this section we shall investigate the validity of equation (6.48). As we saw in the previous section the intermolecular potential can be written as: u(r) = uh(r) + u.(r) , (6.49) where the subscripts h and 8 refer to the ’hard’ and ’soft’ parts of the potential. This simple division assumes the solute-solvent pair potential is a function of separation r only. The excess chemical potential is: Au; = -kan = -kT6n] = -kT!m - kan . (6.51) s and therefore Auz could be split into two factors. The 89 point is that Bh and B. are generally not independent random variables so equation (6.51) is invalid. A proper interpretation can be obtained by writing the average in equation (6.50) as 14: I d?" exp[-pU(?F) -th] exp(-BB.) I d?” expl-BU(;") ] I dr'exp[-BU(;') -th] I df'expl-BU(?') -pah] exp(-BB.) I df'exp[-BU(3') ] I df'exp[-BU(f') — BBh] = h . (6.52) where the subscript h on the second average refers to a conditional average. This means the second average is taken assuming that a hard solute particle already exists in the liquid. The excess chemical potential is then: Au; = -kT6n -kT.€n.h = Au: 4» Apt/h . (6.53) The first term, Au:, is the free energy to add a single, hard solute molecule at a fixed position. The second term but,“ is the conditional free energy to couple the soft part of the solute-solvent interaction given that the hard part of the potential has already been coupled. This is perhaps 90 better understood using the coupling parameter approach. As we saw earlier the excess chemical potential can be found by utilizing a coupling parameter to turn on the potential and essentially grow the particle in the liquid. Since this process is carried out reversibly it does not matter what path we take to couple the particle. One possible way is to turn on the hard part of the potential first and then turn on the soft part, this is exactly what we have done in equation (6.53). (As you may have guessed one could just as well perform the calculation in the opposite order and couple the attractive forces first and then calculate the. conditional averages for the repulsive potential.) The idea of breaking up the salvation process has been utilized by many researchers in developing simple models to predict solubility. The salvation process is normally modeled by the following physical process: (1) One makes in the solvent a cavity just large enough to fit a solute molecule. The free energy associated with this process is called gc". For this part of the process the solute is considered to be a hard sphere. (2) One now allows the solvent and salute to interact with the soft potential; the 52.53, associated free energy is gint Au; = g + g . (6.54) From our previous discussion it is obvious that the cavity * free energy is equivalent to A"): and the interaction free 91 * energy is equivalent to Autnv Equation (6.54) provides the basis for most successful analytic theories to date on solubility (the term successful implies both accuracy and flexibility). We shall now adopt a model, due to Reiss et al.54 and Pierroti,55 in order to analyze our data. The primary assumption made by these authors is: Since there is no analytic expression for gun, in a real liquid one exploits the van der Waals model and calculates gen for an effective hard sphere system whose radii are chosen to best reproduce features of the pure solute and solvent. Although our solute (Xe) qualifies as spherical, the organic solvents certainly are not. One must regard this appoximation as, in some sense, averaging over spatial configurations of the solvent. e If we combine our experimental expression for Auz with equation (6.54) we have: Au; = mum. = g + g . (6.54a) Equation (6.54a) can be interpreted in thermodynamic as well as in statistical mechanical terms. If one calls v2 the partial molar volume of solute atoms in the solvent, v: the molar volume of solute gas, and mg the volume fraction of solution occupied by solute, then we have L = (adv: /v2). Substituting for L in equation (6.54a) gives: g + g + 127mg - n'ronz/v‘z') = 0 . (6.54b) cev int 92 Equation (6.54b) has the following thermodynamic interpretation 3: ‘The free energy for transferring at equilibrium one mole of solute gas into the solution is the zero sum of four terms. The first two terms are for the physical process already described. The third term may be interpreted as the free energy change associated with the entropy of mixing -R6n¢é bentropy increase, free energy decrease) in a real solution. Finally the fourth term is the free energy required for isothermal and reversible compression of the solute from its volume in the gas to itsv volume in solution (entropy decrease, free energy increase). The idea that entropy of mixing depends on volume fraction in real solutions is supported by the work of Flory,56Huggins,57 and Longuet-Higgins.58 93 6.5 The Scaled Particle Theory for Hard Spheres We now must calculate the free energy for introducing a hard sphere solute with diameter a2 into a hard sphere fluid with solvent diameter a1. At present there exists no exact theory even for this simple system , although it can be solved numericallysg. However the scaled particle theory (SPT) has proven to be a very good approximation. Details of this theory are quite lengthy (see refs 60,61) so only a brief outline will be given below. The basic idea of the SPT is to calculate the reversible work to produce a cavity at fixed position in a fluid of hard spheres. A cavity is defined as a sphere of radius, r, from which the centers of all fluid particles are excluded. It is apparent from Figure 13 that a cavity of radius, r, in a fluid of spheres of diameter, a , is produced by the 1 introduction of a hard spherical solute of diameter, a2, such that: r =-—1———£ . (6055) The equivalence of these two processes allows us to conclude that the free energy to create a cavity in the fluid is the same as that to introduce an appropriately chosen hard sphere solute. 94 Figure 13. Definition of a cavity in the SPT. 95 We can see from Figure 13 that for r=a1/2 the diameter of the solute molecule goes to zero. Thus, a cavity with r = a1/2 corresponds to a point solute. If the cavity is smaller than a1/2 only one solvent center can occupy it at any time. The corresponding probability of finding the center of a molecule in the cavity is: p (r) = 4 urap r s a /2 (6 56) l 3 ’ 1 ' where g) is the number density concentration of solvent molecules. The probability that the cavity is empty is therefore: _ _ _‘ _ 4 3 The reversible work theorem is a general theorem of statistical mechanics which states that the probability of observing a fluctuation in a system is equal to the reversible work required to produce the fluctuation divided by kT.62 Thus the probability of finding in the solvent a cavity of radius r that is empty is: p°(r) = exp(-W(r)/kT) . (6.58) Since we imagine a process of empyting the cavity at constant N,V,T, we may identify W(r) may with the change in Helmholtz free energy for the process and write: 96 polr) = exp(-AA(r)/kT) - (6.59) Substituting (6.57) in (6.59) we find: AA(r) = -kan( 1 - g- ur3p) . r 5 81/2 (6.60) In the limit of a macroscopic cavity the free energy is just the thermodynamic work of compression: AA(r) = PV = P g nr3 , r>>a1 (6.61) where P is the macroscopic pressure in the liquid. We now have two exact results for the cavity free energy. In order to bridge the gap between these two limits we introduce a new function, p°(r+dr), the probability that a cavity of radius r+dr is found empty. This may be written as: poir+drl = po(r)po(dr/r) . (6.62) where p°(r) was defined earlier and po(dr/r) is the conditional probability that a spherical shell of width dr will be empty given that the sphere of radius r already is empty. If we expand p°(r+dr) to first order in dr, we obtain: 97 dpo(r) po(r+dr) = po(r) + dr . (6.63) dr Equating (6.62) with (6.63) gives: dpo(r) Poirlpo(dr/r) = po(r) + dr dr dén Po(r) Po(dr/r) - 1 = ‘———————— dr . (6.65) dr We now introduce the auxiliary function, G(r), defined by: p4xrzG(r)dr = 1 - po(dr/r) . (6.66) Substituting into (6.65) we have: dén p°(r) 2 ————————— = -p4nr G(r) , (6-67) dr which upon integration, yields: l' on p°(r) - 0n po(r=0) = -pI41rlzG(l)dl . (6.68) 0 Since po(0) = 1, equation (6.68) reduces to: r 2 0n p°(r) = -pI4ul G(l)dl , (6.69) O 98 or, using equation (6.59): r 2 AA(r) = kTpI4xl cum). . (6.70) 0 The work required to create a cavity of radius, r, is a scaling process, in that we build up the cavity from l=0 to l=r. This is the same as building up a hard sphere particle at a fixed» position in the fluid, hence the name Scaled Particle Theory. 60 To find AA(r) Reiss gt, :5; used statistical considerations to suggest a simple analytic form for G(r): G(r) = A + (B/r) + (C/rz) . (6.71) The coefficients A, B, C, are determined by using the limiting forms of G(r): G(r) = (1 - gnrap)‘1 r s a1/2 G(r) = P/kTp r>>a1 , (6.72) along with the assumption that AA(r) and its first two radial derivatives are continuous at aI/Z. The result is: _ 2 a AA(r) - K0 + Kir + Kzr + Kar , in which the coefficients are: 99 K. = kT{ -6n(1-y) + 3 [y/(1-y)]2} - (nPaia/G) x1 = -(kT/a,){(6y/(1—y)1 + 13[y/(1-y)12} + «9.12 K2 = (kT/612)([12y/(1-y)1 + 13[y,(1_y,]2, - ZuPa1 _ 4 K3 -§7¢P e (6.73) In Eq. (6.73), y=ua13p/6, is called the packing fraction of the hard sphere solvent (it is the fraction of space occupied by the fluid). The salvation free energy, AA(r), is for a constant T,V,N system, however this is the same as the Gibbs free energy in a T,P,N system provided the average volume (V) in the latter is equal to the exact volume V in the former system.14 We can therefore equate AA(r) with the cavity energy, gc.v, defined in the previous section. Figure 14 shows gen/kT as a function of the reduced cavity radius R’= ant/a2 for a hard sphere solvent with packing fraction y = 0.5. The open circles are those described by equation (6.73) while the closed circles are described by the exact relation (6.60). The cavity free energy increases monotonically as a function of cavity radius and is always positive. GCAV (kcal/mol) 20 15 10 100 I I I I I I I I I I I I I I I I I _l ‘1 y=0.5, al=5A _ — 0 SPT —' >< Eq.(6.60) a -— —1 lyvl XL k IXI I I I I I I I I I I I I I I . I 2 4 6 8 _ 10 R12 (angstroms) Figure 14. gcavm‘zi/k’r 1:, cavity radius R“. 101 To summarize, the free energy required to create a cavity large enough to accomodate a hard sphere of diameter a2 in a hard sphere fluid with particle diameter a1 and number density p1 is, by rearranging terms in Eq. (6.73): - - 2 - Sc“,- kT{[6y/(1 y)][2(r12/al) (rm/an] +[18y2/(1-y)2][(r12/61)2 - (rm/a!) + 1/4 1 - mum} I . (6.74) O’HH +uPa3[% (rm/a1)3 - 2(r12/a1) + (rizlail - with r12: (a1+a2)/2 , and y=naaap1/6. ‘The total excess chemical potential is defined by equation (6.54): A"; = gcsv+ gint ° (6°54) An aside for experts: The SPT also leads to an equation of state for the hard sphere fluid which is exact up to the third virial coefficient, which agrees over the entire range of fluid density with the equations of state obtained through machine computation, and which is identical to the equation of state obtained by exact solution of the Percus-Yevick equation.63’35 6.6 Application to Real Liquids The SPT is extended to the study of real liquids by: (a) using the actual liquid density of the solvent and (b) using 102 a simple model and the heat of vaporization to determine the optimal hard sphere diameter a1 of the solvent. This model essentially examines the salvation process for the solvent in its own liquid. The excess chemical potential per male for a single component gas-liquid system is given by equation (3.5): * = -RT , where 06 and p9 are the molar densities of the liquid and gas respectively. Assuming the gas phase is ideal, we have: * An = -RT 6n ( at RT / P9 ) , (6.75) where P9 is the vapor pressure of the liquid. We now apply our model to this system and equate the excess chemical potential to that of cavity formation plus attractive interaction: * Au=~RTaOn(szT/P9)=g +3 . (6.76) cav int It follows from equation (6.76) that: Ah = h 4-)) As csv int' cav int (6.77) H m + W -* The excess molar enthalpy. Ah , can be related to the molar heat of vaporization, AHv as follows: 103 -AH ( 1 - Pv / RT) = AE*+ RT - RTza v 6 L = h + h - RT + RTzd s -AH (6.78) t C. v cav in where the heat of vaporization is defined as, AHV= (H "- s H ), a is the coefficient of thermal expansion for the liquid it pure liquid, and VI. is the molar volume of the liquid. We assume ,as usual, that Pvt/RT << 1, so we can neglect it (see appendix # for a detailed derivation of 6.78). We shall now look at the interaction term. The molar energy of vaporization of a normal liquid is approximately given by:36’64 Q AU;m -2nNApI u(r)g(r)r2dr , (6.79) 0 where u(r) is the pairwise additive intermolecular potential , g(r) is the radial distribution function (see Eq.(6.31) ), and NA is Avogadro’s number. On the other hand, the molar energy of interaction can be derived from equation (6.33): Q “int: 4uprIau(r)g(r)r2dr . (6.80) Hence, it follows that for such a liquid: AU = -u / 2 . (6.81) 104 Now we invoke the thermodynamic relation: H = E + Pv = K + U + pv , (6.82) where E is the total internal energy, and K is the kinetic energy. We may then reexpress the heat of vaporization as: AH = H - “C = (Kq- Kl) + (Ug- Ut) + P(Vq- Vt) a RT + AUv . (6.83) We have assumed that V§>>Y£, K7“ KL’ and the gas is ideal. Plugging (6.81) in (6.83) we find: I1 = 2RT - ZAH . (6.84) int v Once again we use the fact that, in a liquid, the PV term is negligible, hence u‘n{= h . Finally, by introducing equation (6.84) into (6.78) we get the following relation for the standard molar enthalpy of vaporization: AH =h +6:t RT2+RT . (6.85) V CIV This is a general equation which should be satisfied by any theory which predicts cavity terms. Now we can use this to determine the effective hard sphere diameter in the SPT. The enthalpy of cavity formation is found using standard 105 thermodynamics: i.e. he": -T2(a(gc"/T)/8T)P 55: hm: a, 672 (y/z) [(6/2) + (say/22) + 11 . (6.86) where z=(1-y), y is the packing fraction, and “L is the thermal expansivity of the liquid. We have also assumed that (1/v)(av/6T) = -(l/y)(ay/8T). Putting this result in (6.85) we find, finally: 2 AB = RT + a, RT2[ (115%) . (6.87) ' (l-y) The SPT is related to real fluids by using the heat of vaporization, density, and thermal expansivity to find an effective hard sphere diameter for the fluid. It is by no means a first principles theory since it uses the experimentally measured values for these variables (there is no theory to date that can predict the density of even atomic liquids let alone molecular liquids). However, the SPT is consistent with our current goal of being able to predict solubility by'knowing only bulk properties of the solute and solvent. Other techniques exist to determine the effective hard sphere diameter of the solvent, a‘, but they require a detailed knowledge of the intermolecular potential and lots of computer simulation (or very simpLe molecular 43,49 structure). This option was not presently available to us. 106 6.7 Evaluation of the Cavity Term: Table IV lists the heat of vaporization, number density, and thermal expansivity for each of the solvents studied. Since our experiments were generally done in a range of 5-50°C we have chosen an intermediate temperature of 25.0°C to evaluate a‘. In theory, a1 should be temperature independent. In order to evaluate the number density we require the mass density29’30’65 and the molecular weight, i.e.; p'(mol/cm3) = pu(g/cm3)/M.W.(g/mol). Table V and Figure 15 show the calculatated values of a1 for the solvents at 25.0°C, we have not included nonadecane, eicosane, dodecanol, and tetradecanol because they are normally solids at this temperature. The values of ai were calculated from solving equation (6.87) for y and using the known density to solve for a1 by the relation; a1=(6y/xp1)“3. ’The points on Figure 15 vary monotonically in each homologous series with number of carbons but they, bear no simple relation to actual chain lenghts. In alkanes, for example, the C-C bond length is about 1.5A. Of the hydrocarbon solvents we studied, the more polar molecules tend to have a larger effective radius than their nonpolar analogues. One can obtain values of a1 other ways besides the one we use here. For example, one can use gas viscosity,66’67 or second virial coefficient data.68’69 However, for all but a few of the short alkanes (and for those) molecules the agreement with our 'values of a1 is 107 good55) the results are limited by lack of available data. In order to evaluate go" we use Eq. (6.74). For the cavity radius we have rm= (a1+ a2)/ 2. For Xe we took the hard core molecular diameter to be a2= 3.973A; this is the potential parameter a (or 21161.0) in the Lennard-Jones (6-12) potential for Xe.73 The calculated values of go" are in Table V and Figure 16a. All the values of go" are positive and range from 2.5 to 9.4 kcal/mole. This shows that for a hard sphere fluid it always takes a positive amount of work to make a cavity. The fact that go" is positive means that its contribution will always tend to lower solubility (L=e-Au/"), therefore, systems with low solubility are dominated by repulsive interactions (as intuitively expected). In contrast, the experimental quantities Au; for these solvents, with the exception of formic acid, are negative and in the range from about -0.4 to -l.0 kcal/mole. Since, by equation (6.54), Au; is just the sum of gen and 81:". these results imply that g1M must be negative and its magnitude must be a few to several kcal/mole. This isn’t surpring since one would Table IV. Heat of Vaporization, 108 number density, and thermal expansivity for organic solvents at 25.0°C. alkanes n-C AHV(::f:) pntxtogéclattso-gc) 5 6.39 5.19 1.558 6 7.54 4.5788 1.368 7 8.74 4.0852 1.266 8 9.92 3.6851 1.183 9 11.10 3.3535 1.114 10 12.28 3.0740 1.071 11 13.46 2.8399 0.9704 I—12 14.64 2.6378 0.9417 13 15.83 2.4613 0.9432 14 17.01 2.3083 0.9300 15 18.20 2.1706 0.9097 16 19.22 2.0524 0.8986 17 20.6 1.9409 0.879 18 21.7 '1.8434 0.8661 cycloalkanes n-C AHV(::::) pntxiogécIat(10-3C) 5 6.85 6.1644 1.3298 6 7.90 5.5400 1.2475 8 10.36 4.4702 0.98805 109 Table IV. cont ..... alkanols n-C AHV(::?:) pn(x10§écIa£ 1 8.94 14.79 1.189 2 10.18 10.267 1.093 3 11.51 8.0178 1.006 4 12.50 6.5509 0.9265 5 -13.61 5.5443 0.8804 6 15.00 4.8128 0.8720 7 16.20 4.2446 0.8828 8 17.00 3.8043 0.8491 9 18.60 3.4444 0.7875 10 19.82 3.1453 0.8283 11 21.00 2.8988 0.8105 carboxylic acids n-C AHV(::::) pn 1 11.03 15.891 1.0249 2 12.49 10.475 1.0114 3 13.7 8.0341 1.0974 4 15.20 6.5163 1.0693 5 16.56 5.5117 0.99941 7 18.1 4.2244 0.91035 110 Table IV. cont ..... alkanals (aldehydes) n-C AH (kcal) p (xiogécIa (io-3c> v mole n 3 3 7.09 8.2083 1.4619 4 8.05 6.6543 1.3084 5 9.17 5.6298 1.2474 7 11.40 4.2907 1.0333 perfluoroalkanes n-C AH (kcal) p (xioiéCIG (iO-3C) v mole n 5 6 7.606 2.9844 1.6698 7 8.69 2.6978 1.5047 8 9.77 2.4216 1.4119 Hard Sphere Diameter, a1 (A) 111 10 ’- I I r I I I I I f I I I I I I I l T I r _ 9 __ .2 .. o _ _ o _ _ o a _ o - — O —-( 8 c o - _. )1 O .. .— X )1 O —( __ X __ 7 _ )1 O .. - X E o O Alkanes ‘ I- -( - 31 5 ° )1 Alkanols 7 6 — a O , — - Cl Ac1ds ~ - E 6 <> 9 .— .l _ B + o 0 Cyclo 4 5 — X Perfluoro — r- .g + '7 - + Aldehydes 7 4 fl 1 I l I l L I l I l I i I I I L 1 d O 5 10 15 Number of Carbon Atoms Figure 15 Effective hard sphere diameter for 45 organic solvents 1; number of carbon atoms. 112 Table V. SPT calculations for organic solvents. The second column is the effective hard sphere radius. The third column is the packing fraction. TEEI£ourth is the cavity free energy for Xe. alkanes n-C a‘(A) y gégei/mou 5 5.476 0.44614 3052 I 6- 5.877 0.48656 I 3396 7 6.226 0.51611 3645 8 6.543 0.54055 3861 9 6.838 0.56143 4052 10 7.108- 0.57791 4191 11 7.386 0.59925 4463 12 7.623 0.61176 4577 13 7.836 0.62006 4613 14 8.044 0.62914 4682 15 8.251 0.63838 4765 16 8.436 0.64516 4805 17 8.635 0.65423 4907 18 8.813 0.66075 4958 cycloalkanes n-C a1(A) y gcéeal/mol) 5 5.286 0.47681 3766 6 5.582 0.50450 4019 8 6.233 0.56678 4775 113 Table V. cont.... alkanols n-C a1(A) y gcé$°"m°" 1 4.082 0.52668 7366 2 4.686 0.55305 6854 3 5.156 0.57554 6703 4 5.580 0.59588 6667 5 5.948 0.61083 6595 6 6.274 0.62238 6508 7 6.567 0.62929 6327 8 6.843 0.63822 6274 9 7.134 0.65467 6521 10 7.358 0.65612 6279 11 7.591 0.66382 6289 carboxylic acids n-C a‘(A) y gcésaL/moi) 1 4.092 0.57015 9429 2 4.746 0.58623 8143 3 5.189 0.58790 7131 4 5.610 0.60244 6872 5 5.986 0.61891 6854 7 6.606 0.63774 6601 114 Table V. cont.... alkanals (aldehydes) n-C a1(A) y gcéeal/mol> 3 4.780 .46945 4210 4 5.239 .50109 4337 5 5.623 .52417 4414 7 .6.342 .57316 4821 perfluoroalkanes n-C a‘(A) y gcésai/moi> 6 6.662 0.46209 2520 7 7.045 0.49389 2740 8 7.417 0.51733 2876 115 10000 i T r T I 7 I I l I 7 fl 1 1 I ‘3 1 L T = 25.00 C O Alkanes ‘ 8000 .. C) ‘1 Alkanols « )— - C1 Acids _ A __ )1 -l 0) C1 0 Cyclo 4 3 h K 11 El] 0 .4 8 ~ :1 11 g r! X Perfluoro' )1 I! I! -( E 6000 r t Aldehydes -— 3 L -( z ' + o O C' '3 d . o _ QB ‘ + + t O O O 0 q 4000 L— o O o 0 _ o b- 0 d O o . h K x d L X . 2000 ' . . - 1- 1 . 1 | . .. . I . . .. O 5 10 15 20 Number of Carbon Atoms 500 ,_ a II I 1 I l 1 I I l I I I l I 7 I 1 . 4 Z i 250 :— O Alkanes -j “' K Alkanols 7 2 ” C1 Acids . E Q t 0 Cyclo 1‘ > t X Perfluoro < (0 " .7 3 -250 :- 0 f Aldehydes J - - -( IN _ x x -( 1 )— x X n -( Q -500 "" ’2‘ xx 1‘ x “ .- E n n x X -( .. c] C] E] ' O -l .. + + O O O .4 _ o o o 4 -750 ;- o o O o 1 - 0 . .. 9 0 .. - 2L 1 l L l I l l L l 1 I l l l L -1000 . 9 L 0 5 10 15 20 Number of Carbon Atoms Figure 16. (a) g“W for Xe in 45 organic solvents at 25.0°C. (b) corresponding excess chemical potential, (in: = -RT&nL.. Molecular Diameter 02 (A) 116 1"— _ O- LiLlllLll 111 1 l 1 ' Lll'ill"‘.ll] 0 l 2 - 3 4 5 6 Polarizabiiity a (10'24 cma/molecule) Figure 17. Molecular diameter ¥§_polarizability for several inert gas solutes. Solid line shows extrapolation for noble gases. -—R'I‘ [u(L) = Ape” (kJ / mole) 117 O Tetradecane -* U Decane 0 Hexane 0‘ O I Ln '1,[ ..l I 111' "lll‘III fl C1 _.I d .1 d _1 q O 1 2 :3 4 Polarizability a (10'24 c-ma/molecule) Figure 18. Excess chemical potential 11 solute polarizability for noble gases in Tetradecane, Decane, and Hexane. Filled in points along ordinate are the SPT predicted values for a hard sphere solute of 2.58A. -—R‘I‘ ln(L) = Ape” (cal / mole) 4000 . ei,, , r, r k t l ' T T ’ ' 7 t. T = ZSIP C ‘ 3000 — __‘ '_' 1 U Methanol ~ 2000 0 Ethanol — 118 -24 Polarizability a (10 cma/molecule) Figure 19. Excess chemical potential 1; solute polarizability for noble gases in polar solvents methanol, and ethanol. Filled in points along ordinate are the SPT predicted values for a hard sphere solute of 2.58A. 119 intuitively expect the attractive interaction to lower the free energy of solvation and increase the solubility. Figure 16b shows the experimental values of Au; for Xe in the alkanes; it is encouraging to note that the SPT has much of the same systematic behavior. The physical significance of gc.v is supported by extrapolation of measured solubilities of inert gases in 55’66’67 In this technique the excess chemical solvents. potential is plotted versus solute polarizability, up, in a single solvent at fixed temperature. It has been shown that extrapolation of this data to zero polarizabilty is equivalent to finding the free energy required to introduce a hard sphere of diameter 2.58A into the real solvent.55’72 The hard core diameter is determined by extrapolating a plot of solute diameter versus polarizabilty. Such a curve is shown in Figure 17 for the inert gases He through. Xe. Figure 18 shows a plot of -RTlml. versus up for the inert gases in n-hexane. Extrapolating to ap=0 is equivalent to writing: * . Auz (ap= 0, a2- 2.58A,T) - gc.v(a1,az-2.58A,91,T) . (6.88) This value can then be compared to the SPT prediction of equation (6.74) for a solute of diameter 2.5&A. As Figure 18 shows, the predicted gcw agrees to within 2% of the extrapolated value. Pierotti has used this test on a wide 120 variety of non-polar solvents with good results.55 ‘We performed this test for some of our solvents in order to see if the SPT was applicable. Even for alkanes as long as n-Cunao the calculated gc" agrees with the extrapolated value to within 5x ( we couldn’t test most of our solvents in this manner due to lack of experimental data). The case is different for the polar solvents. Figures 19a and 19b show the extrapolated and SPT results for methanol and ethanol. Clearly for these solvents there is a discrepancy. This can be traced back to our derivation of the effective hard sphere diameter from the solvent heat of vaporization. We recall that Eq. (6.79): 0 AUva -2nNApI'u(r)g(r)r3dr , (6.79) o . holds for normal liquids in which there are only two -body radial forces. For polar liquids one must account for many-body correlations. This renders equation (6.79), and therefore equation (6.87): 2 AB = RT + a:z RT2[ 13131-3?) , (6.87) ' (l-y) invalid. We can also see that for methanol and ethanol the descrepancy between the two results decreases as the solvent chain length increases. Although we can’t test it at this time; we are confident that for the longer acids, alkanals, 121 and alkanals the SPT works much better. It should also be mentioned that for the highly polar solvents the effective hard sphere diameter evaluated from solubility data together with the SPT agrees very well with other experiments; this involves using the SPT relation for gc‘v to evaluate a1 from the extrapolated result.72’73 In other words we use a plot like Figure 19 to find the free energy to dissolve a hard sphere of 2.58A. The SPT is then applied to this result to give the effective radius of the solvent a1. This has proven to be a powerful method for estimating the molecular diameter of a solvent. Such a technique has even been applied successfully to water, which is a very complex 11gu1d.7°’73 From the calculated values of gc"(T), we obtain the enthalpy of cavity formation by the Gibbs-Helmholtz equation: h = -'r"’[a(gc"/ T)/ 8T] . (6.89) OIV P In this calculation .we again neglect the temperature 54,55 dependence of a‘. Using our previous expression for g (T), we find: IV C 36y (1-y) he": 31.2% Téi [13; [A] + lel + 1] . where we have used the notation, 122 [A] = [ 2(r12/ a1)2 - (r,2/ a1) 1 .and [B] = [ (r12/ a1)2 - (r12/ a!) + 1/4 1 . (6.90) This expression reduces to equation (6.86) in the limit a =a (i.e. r /a =1). Figure 20a shows h (25°C) versus 1 2 12 1 cav for our solvents. For comparison Figure 20b shows carbon -1: the experimental excess enthalpy of salvation Ahz for these solvents. One may obtain the entropies of cavity formation, s (25°C), from g and h by the thermodynamic relation ClV CIV 08V s = (h ' 8 )/ T. The 3 values thus obtained are CIV CIV CIV 0" shown on Figure 21a. However, they appear to be unreliable. They do not have the systematic dependence with series that one might expect, nor do they correlate with our experimental results (Figure 21b). In order to check a possible cause for this we have plotted in Figures 22-24, respectively, the heats of vaporization (ARV), number density (pa), and thermal expansivity (5.2) versus carbon number for our solvents. Of these three input parameters in the SPT only the thermal expansivity seems to vary roughly in some of the homologous series. Perhaps better he“, (cal/moi) AHz'. (cal/moi) 123 10000 I . . . eI . , " i __ O Alkanes : ' ’1 Alkanols # —- Cl _. 8000 __ U 0 Cl Acids 4 - D 0 Cyclo i _ x n n x K :x n :x Perfluoro . 6000 — t Aldehydes q a -l - O o .4 .. O O _ t * ‘ 0 O o O 0 j o 4000 - 0 0 ~— ... o O '1 - x .l - X X .4 L- I -4 2000 #4 l l l l l J 1 J 1 I. l L 9 l 0 5 10 lo 20 Number of Carbon Atoms L fir I r T I Y I f I T 1' d t O Alkanes * r K Alkanols ‘ - 1000 - o Ac1ds — L ocydo ~ ” x x X Perfluoro ‘ t x t Aldehydes * -1500r- —* - # . " O + 1 9 e o P ‘1 n g x z 1“ O O O " -2ooo+— C" o O o —. l- 0 O O 4 I o i - — —1 2500 l L f£ 1 l J; l L 1 l l :l_ I ‘ L l l 0 5 10 15 20 Figure 20. (a) h CQV Number of Carbon Atoms for Xe in 45 organic solvents 11 number of carbon atoms.(b) corresponding 1: experimental excess enthalpy, Ahz . ASZ', (cal/moi K°) 124 4 I I I I I I I I I I I I T I I I I I I q _ 0 Alkanes : .— D .I 3 -— a x Alkanols — : O U Acids : 6‘ - . a: 2 — 0 Cyclo — s 2 * E : xx 0 + x D X Perfluoro 1 § 1 _0 X x + Aldehydes — 3 : =1 + 9 o o o . n . 3 " O O K O O “ 8 O _ )1 o I! O o O _ U) h + n O " I- )1 ' O '1 — K n -1 _1I-— X —— : o _2 l l l l J l l l l I 1 l l l I L l l l O 5 10 15 20 Number of Carbon Atoms _2_5 I I I I I I I I I I I I I I I I I I I x 3 I x : -3.5 :- D + + _q -0 + . _ o + . __ c: . 0 ° . -4 0 _ O O O O O — I 0 o . .. D o x x ‘ -4.5 :" n n x x O Alkanes i .. F x " x I! Alkanols ‘ : x 0 Acids : -50 -- 0 Cycle m : X Perfluoro : ~ 1 + Alldehydes ~ P _5.5 L l l L ¢ 1 l 1 l l l 1 1 l l l l O 5 10 15 20 Number of Carbon Atoms Figure 21. (a) 3mV for Xe in 45 organic solvents 11 number of carbon atoms. (b) corresponding * experimental excess entropy, A3,- Heat of Vaporization, AHV (keel/mole) 25 20 15 10 125 + . 8 l O x C] o x + l 5 10 O -4 O --l O i o s O .1 Alkanes ‘ Alkanols ‘ Acids - Cyclo 1 Perfluoro -* Aldehydes ‘ l I l 1 l l 1 5 20 Number of Carbon Atoms Figure 22. Heat of vaporization for 45 organic solvents at 25 C 11 number of carbon atoms. Density, ,0n (102‘ molecules/CHIS) 126 Li I I fr I fiI I I I r I I I r I 15 "-11 —' ' T = 25.0° C O Alkanes ~ “' ’1 Alkanols « ' U Acids a ' g 0 Cyclo 4 10 '— X Perfluoro ‘ - t Aldehydes ‘ .. a . .. a i _ 0 p o 5 — 5 5 o -— .. 5 1 5 6 8 I— X X g .I X ‘3 O O O .— O O O -I b ' , . . L I L . 1 1 L . 4 O J l l l l J 0 5 10 15 20 Number of Carbon Atoms Figurg 23. Number Density for 45 organic solvents at 25 C 11 number of carbon atoms. 'l'hermal Expansivity, (XI (l/Co) 2.00 1.75 1.50 1.25 1.00 127 .— I I I I I I I I ‘1 " O Alkanes ._‘11 " x K Alkanols 1 l- ... - c1 Acids 4 _ o g _ X ' _ ¢ 0 Cycio —_I ‘- 0 X X Perfluoro j ' o . - ‘* O ‘- Aldehydes 1 I_ + o —n" - O . _ :1 Cl :1 + o . - 11 x g O O . d — x u 3 -I _ z x _q ' :i — 4 l I ' ' 1° l ' I 1 I l J1 0 5 10 15 ‘20 Number of Carbon Atoms Figure 24. Thegmal Expansivity for 45 organic solvents at 25 C 11 number of carbon atoms. Table VI. Cavity enthalpy and entropy. 128 alkanes 5 3251 0.67 6 3482 0.29 7 3710 0.22 8 3900 0.13 9 4067 0.05 10 4224 0.11 11 4332 -0.44 12 4467 -0.37 13 4613 -0.002 14 4739 0.19 15 4850 0.28 16 4929 0.41 17 5067 0.54 18 5148 0.64 cycloalkanes n-C cav(§£§) scav(S£%kl 5 3725 -0.14 6 3980 -0.13 8 4398 -1.26 129 Table VI cont.... alkanols n-c ...(fi - l 7760 1.32 2 7027 0.58 3 6675 -0.10 4 6440 -0.76 5 6291 -1.02 6 6336 -0.58 7 .6335 0.03 8 6190 -0.28 9 6289 -0.78 10 6366 0.29 11 6385 0.32 carboxylic acids -C g, 3.9}. . n cav mol. cav mall: I 9730 1.01 2 8538 1.33 3 8046 3.07 4 7820 3.18 5 7620 2.57 7 7017 1.39 130 Table VI cont..... alkanals (aldehydes) -c 21'.) 32'. ‘ n cav mot cav moLK‘ 3 4563 1.18 4 4510 0.58 5 4611 0.66 7 4722 -0.33 perfluoroalkanes _C 2}. L“. . n cav moi. cav mot]:J 6 2893 1.25 7 3046 1.03 8 3166 0.97 131 measurements of these are required. In any case it appears as though the son’s are approximately centered on zero, whereas the total excess entropy, As: , is about -4 cal/mol it. We conclude that the SPT alone does not adequatly predict the salvation entropies. This is in sharp contrast to previous workers, who used the mole fraction scale to analyze their results.55’67’7o’71’74 A surprising result of these SPT calculations is that the cavity formation process is almost completely enthalpic (energetic). Table VI clearly shows that the entropic term is less than 10% of the enthalpic term (this is even more enthalpic than our experimental data; recall Ah:/TA§:81.8). This is in strong contrast to a real hard sphere fluid which is completely controlled by entropic considerations. The key difference between these two situations can be seen in equation (6.90): 36y (l-y) hm: RTza‘c 1%; [If—y [A] + 2[13] + 1] . - (6.90) If the fluid were truly hard spheres then the thermal expansivity and hence the cavity enthalpy would be zero. By putting in the real solvent behavior we are accounting for the full interaction between solvent molecules. At first this may seem improper, but if we write out explicitly the expression for the excess chemical potential due to a repulsive interactions Aph from equations (6.52) and (6.53) 132 we find: A“: = _kT m jd¥"exp[-BU(?~") -thl . h fd?"exp[-BU(;')] It is clear that in evaluating go" we should include the full solvent-solvent interaction. It is only the salute-solvent interaction which is broken up. Whether the SPT properly accounts for this by using the real expansivity requires further testing. To complete our present discussion Figure 25 shows the calculated values of gc“(T) for six typical solvents, one from each homologous series. These results were obtained by applying equations (6.74) and (6.87) at each temperature for which the calculation is made; i.e., a1(T) values were obtained from equation (6.87) for each temperature. The AH"p(T) data for these calculations were obtained from Watson’s relation and critical temperature (Tc) data 33: 0.38 AHV(T2) = AHV(T‘) [(1- T2/ Tc) / (1- T1/ T9] (6.91) The calculated temperature dependence of gc"(T) on Figure 25 is in the opposite direction from the observed values on Figure 9. However the slopes of the lines on Figure 25 are not the entropies of cavity formation. a... (cal/Incl) Cavity Energy. Figure 25. Cavity free energy, 9 133 CGV in 6 representative organic solvents. 8000 7000 6000 5000 4000 3000 2000 1211' £01: I II I II [ill I II III I‘II I II I II "" u-I :__ n-c,H,coon __': _"__ tremor! .. ..J — a - X~ n-C3H7CH0 - P - — n u d I- n x a: .- ' “‘CJHI ‘ - d : NH n-CJ' 1‘ : _1_ll I [II [I I II I II I 11.1 In I ll 10 20 30 4O 50 . 0 Temperature. C Xe 134 6.8 The Interaction Term After the cavity is formed in the hard sphere solvent, one introduces a solute molecule into the cavity and lets it interact with the solvent molecules; the free energy associated with this step is glut. The natural next step is to calculate g1M for Xe in our solvents. This proves to be very difficult even for our simplest solvents. Since zinc is a Gibbs free energy one may write it in terms of enthalpic and entropic contributions as: glut: hint- Tslnt m uint - Tsint ' (6°92) This means that the two principal quantities which determine giM are the potential energy of interaction, ulnt’ and the entropy of interaction, sint. Both of these are difficult to calculate. Solvent reordering, caused by the solute-solvent interaction, is probably the origin of most of the entropy of interaction and one therefore expects that sun} 0.54 Neff and McQuarrie75 include a term in their expression for gun which may account for this entropy; it gives the contribution to the free energy (due to changes in solvent-solvent structure when a solute is introduced. This term [equation (27) of Ref. 74] depends on a derivative of the solvent-solvent distribution function, g11(r), which is hard to calculate. 55,71 neglect sint and make the Other workers 135 approximation that gins uint. I believe that sun cannot be neglected, but rather may be commensurate with the entropy of salvation As:. This contention is supported by consideration of the entropy of interaction in pure solvents for which it can sometimes be estimated. For example, Reiss at al.54 have calculated the heat of vaporization for liquid Ar at its boiling point using the SPT and thermodynamics. From their results [equations (4.3)-(4.11) and Table II of Ref. 54] one finds sin£= -4.6 cal/mol K. We have made a similiar calculation for other pure liquids and find sing: -10 cal/mol K for CCl‘ at 25°C, and sun: -12 cal/mol K for n--CGHH at 25°C. The entropy of interaction is found by combining equations (2.5), (6.54), (6.84), and (6.92): * A" = -RTm(pAC/pq) = gcav + uint - Tsint = g + 2(RT - AH ) - Ts cav v int sint= [RT&n(pz/ag) + gcav+ 2(RT - AHV)] / T . (6.93) These sint’s must properly be compared to the salvation entropies for the pure solvent, A§*,which one finds as follows: for the pure solvent we plot the excess chemical potential, Ana“, versus T along the gas-liquid coexistence line. The temperature derivative is then written as: (Inf/aw); (emf/aw), + (adu*/8P)T(dP/GT)O . (6.94) 136 where the subscript c means the derivative along the coexistence line. We then utilize the thermodynamic identity; vi: Ufifi/BP)T , and equation (6.94) reduces to: a s * (dAu /dT)c= -As + Av (dP/dT)c . (6.95) In Eq. (6.95) Av* is the molar volume of salvation, which is approximately equal to the molar volume of the pure liquid.16 Values we obtained for the corresponding entropy defined by equation (6.95) are As*(CCl‘) = -9.1 cal/mol K, and As*(n-CSH“) = -10.6 cal/mol K., both at 25°C. For these two solvents we calculated scfl’s of about 2 cal/mol K. Thus for pure solvents it appears that s1M makes an important contribution to the total entropy. The interaction energy for a single solute molecule with the solvent may be written as: ... a . 9 uint- pLI g12(r) u12(r)dr , . (6.96) where the terms are as defined previously. Evaluation of equation (6.96) is difficult for nonspherical solvents. An approximate method which assumes spherical symmetry has been 55’71 In. this technique a used by’ Pierotti and. others. Lennard-Jones interaction potential u12(r) is assumed for u12(;) and the solvent distribution is taken to be uniform, i.e., 812:1’ outside the cavity radius r12. The integration is over the solute volume, i.e., from r12 to infinity.This 137 is equivalent to an effective Lennard Jones potential with a hard sphere cuttoff at 012: (01+az)/2, as illustrated in Figure 26 (this is a more reasonable cutoff than ro because the actual potential is much steeper for ralkan00)=161¢€12p18:2 [(INTI) - (1/81) ], where the bracketed terms are exactly the same as for the . _ 1/2 _ . . pure fluid, £12— (5182) , and 812' (a1+ a2)/2 . Combining this result with equation (6.102) and (6.99) we find: In u_ t(Xe-’alkaneo) = (Ci/€2)°'5(a12/a1)3[2RT - ZAHV]. (6.106) Table VIII shows the values of‘umt for Xe in the alkanes. 6.9 Comparison with Experimental Data We may now combine our calculated values for he and IV lunt to find the total enthalpy (energy) of solvation. Figure 28 shows a plot of the combined theoretical term, h.= lunv+ u‘n‘, and the experimentally determined enthalpy, A52. versus number of carbons, for Xe in alkanes ranging from hexane (Cs) t0“ hexadecane (016). The theoretical and experimental data both have the same systematic tendency; i.e. the enthalpy becomes less negative for the longer chain alkanes. According to our model this means that the solvent size dependence, a1, in the cavity term slightly dominates over the solvent energy dependence, evin the interaction term. The general agreement between the two sets of data is within 20%. This points out one of the biggest problems in calculations of this sort, viz;to describe the total salvation process we usually calculate Enthalpy of Solvation (cal/mol) 147 O l I I I l I I I I I I I I _500 T htheory=hcav+hint —: _ htheory (8:1) fl —1000 _ -—-‘ ' o o o 0 ° 0 0 o o o o ‘ : htheory (g=gv.w.) : -1500 A x x x X — - x x x x I Z x x X 0 O o o 0 i —2000 _ O O O -—J 7' O 0 exp _ : 0 Z —2500 _ l 1 1 1 1 I l 1 1 1 ' I 1 1 l 1 5 10 15 20 n—carbon Figure 28. Calculated and experimental excess enthalpies of salvation for Xe in alkanes. 148 two separate processes which each have characteristic energies that are opposite in sign and of larger magnitude than their sum. This puts an extra burden of accuracy in the calculation of each term. For example the discrepancy between h. and Ah: for decane is about 16% of the experimental value which in turn is only 7% of he" or 5% of uh“. To claim that a model as simple as ours could do better than this would be an exaggeration. For completeness Figure 28 also includes the theoretical predictions that one would obtain by using a random fluid distribution, g11= g = 12 1, throughout the entire calculation of 11 Agreement int obtained in this case is not as good as that obtained with the hard sphere distribution of Verlet and Weiss.81 We expect that equation (6.106) should not just be applicable to Xe, but should hold for any non-polar solute with a well defined diameter, a2. In order to test this hypothesis we have looked at solubility data for the inert gases He, Ne, Ar, and Kr in several alkanes. Solubility data was obtained from reference (24). The original data were quoted on the mole fraction scale. In order to convert data to the number density (Ostwald solubilty) scale we inverted equation (3.20): L(T) = RT / [Pv1(T)[(1/x2(T)) - 11] , (6.107) where v1 is the molar volume of the solvent and P is the pressure of the gas (all measurements were at 1 atm). Once 149 the temperature dependent solubilities were calculated, we could easily find the excess chemical potential, enthalpy, and entropy using standard thermodynamics . Table IX gives 4: -* -* the values of Apz, Ahz’ and Asz for the inert gases in the solvents hexane, octane, nonane, decane, dodecane, and tetradecane. Figures 29a and bshow the predicted and experimental enthalpies verses 6.01““ for the inert gases in octane and tetradecane, respectively. The solute 77,78,79 parameters are shown below. solute ezlk (oK) az= 82(A) He 6703_-___—276§——__- Ne 34.9 2.78 Ar 119.3 3.448 Kr 172.7 3.59 Table X lists all the relevant theoretical quantities that we can calculate for the alkanes; i.e. g , ii , s , and GOV CIV GOV uint' Once again our simple theory seems to capture some-of the general systematics but is not sufficiently accurate to make reliable predictions. A particular flaw can be seen in the predicted values of enthalpy for the less polarizable solutes like Ar, Ne, and He. Experimental data indicates that for these solutes the enthalpy increases with solvent size, however the theory predicts just the opposite. This 150 Table IX. Excess chemical potential, enthalpy and entropy for He, Ne, Ar, and Kr in the alkanes, including: hexane, octane, nonane, dodecane,and tetradecane. Helium (20 C) n-C A633) Ah:(§%t-) :(;:{K) 6 1801 2271 1.602 8 1975 2312 1.150 9 2048 2715 2.274 10 2102 2260 0.537 12 2214 2153 -0.2076 14 2279 1840 -1.496 Neon (20 C) M 'Mz‘i‘ii" Abbi-3:) Ashfia 6 1587. 1699 0.383 8 1732 2046 1.070 9 1816 1902 0.295 10 1857 1941 0.288 12 2014 2096 0.283 14 2048 1838 -0.715 151 Table IX. cont ..... Argon (20 C) n-c 41.3%) Ah:(§§-{-) 432133;“) 6 436 -297 -2.501 a 594 301 -o.999 9 631 43.9 -2.002 10 676 53.4 -2.123 12 756 276 -1.636 14 811 79.0 —2.495 Krypton (20 C) n-C Au: (:3) A1133) As: (fix) 6 -168 -787 -2.11 a -49.3 -816 -2.62 9 2.86 -716 -2.46 10 44.5 -767 -2.77 12 110 -594 -2.40 14 146 -850 -3.41 152 Table X. 9““. cav, sec“. and “1m for He Ne, Ar, and Kr in the alkanes, including pentane through hexadecane Helium (25 C) M 9...<.%E> ...fi) ...—.9... 2%) 5 1766 1744 -0.07 -529 6 1965 1857 -0.36 -564 7 2109 . 1971 -0.47 -595 m8 -w-~;;55-« 2065 -0.57 -620 9 2345 2147 -0.66 -644 10 2428 2227 -0.67 -664 11 2580 2273 -1.03 -682 12 2647 2340 -1.03 -706 13 2672 2417 -0.78 -714 14 2715 2482 -0.78 -729 15 2765 2538 -0.76 -742 16 2791 2579 -0.71 -752 153 Table X. cont ..... Neon (25 C) 5 1891 1888 -0.01 -1344 6 2103 2011 -0.31 -1430 7 2258 2136 -0.41 -1506 8 2392 2239 -0.51 -1566 9 2510 2330 -0.61 -1622 10 2598 2416 -0.61 -1671 11 2762 2469 -0.98 -1716 12 2834 2542 -0.98 -1774 13 2860 2625 -0.79 -1794 14 2905 2696 -0.70 -1828 15 2958- 2757 -0.67 -1861 16 2986 2802 -0.62 -1885 \‘ 154 Table X. cont... Argon (25 C) "-0 9... SEE-D ..Jfi’ ... 373%; ...(EZi’ 5 2504 2603 0.33 -3137 6 2766 2762 -1.05 -3304 7 2990 2961 -0.10 -3450 6 ' 3167 3110 -0.19 —3563 9 3323 3240 -0.28 -3670 10 3439 3363 -0.25 -3760 11 3659 3445 -0.72 -3840 12 3753 3550 -0.66 -3955 13 3784 3666 -o.40 -3963 14 3842 3766 ~0.26 ~4046 15 3911 3853 -0.19 -4105 16 3945 3915 -0.10 -4144 155 Table X. cont ...... Krypton (25 C) n-C gcav 51%) cos/(2%) scav 76:31:) 1751(3) 5 2646 2771 0.42 -3958 6 2945. 2964 0.06 -4160 7 3161 3154 -0.02 -4337 8 3347 3314 -0.11 -4472 9 3512 3454 -0.20 -4601 10 3634 3586 -0.16 -4708 11 3868 3674 -0.65 -4803 12 3967 3787 -0.60 -4943 13 4000 3911 -0.30 -4976 14 4060 4017 —0.14 -5050 15 4133 4111 -0.07 -5121 16 - 4£;;—~—-4177 -0.03 -5167 Ahz'. (cal/moi) (cal/mol) I 2 0 Ah :X. . I 1 r I . I r - 2000 F“ x E0 OCTANE 1000 _ X experiment — : o O theory ‘ 1 - 4 _ X J O "" —( r J : O '1 L . i -1000-— -d I- O -4 Z 1 -2000 — 4 L 1 i I v I l l L L l l l 41 Li 1 ‘ d 0 50 100 150 200 250 E solute/l k (CK) " . fi fi T I Yfi ' ( Y i 2000 g x _‘1 Z TETRADECANE I 1000 )— O X experiment J : . O theory : 0 :— " i I O i . . 1 —1000 — o —J( _. 4f ' 7 I 1 -2000 _' "‘2 L 7 (.4 ' LQ { L l L J l l l l l l L i l l 1 A 4 0 50 100 150 200 250 E solute/k (CK) Figure 29. Calculated and experimental excess enthalpies of salvation for He, Ne, Ar, Kr, and Xe in (a) octane, (b) tetradecane. 157 seems to indicate that for these salutes we are overemphasizing the attractive contribution to h . A I possible expanation for this lies in our use of the single fluid approximation for g i.e. g11(r/a) = g12(r/a12). It 12‘ has been shown that this approximation gets worse as the solute and solvent become more disparate in size. The most prominent effect is that the main peak in the correlation function tends to become smaller as the ratio of salute to solvent size decreases.83 An interesting feature in Figure 29 is that the enthalpy increases as the solute becomes less polarizable, and crosses over to a positive value at argon. If we recall equation (5.4): A)?" 9.1: = 1,..32 , (5.4) dT a T we see that for the less polarizable gases the Ostwald solubility actually increases with temperature. This hs apposite to our usual experience (when you heat a liquid you drive it towards the vapor phase!), but is easily understood from our model, i.e. as the solute becomes less polarizable it behaves more like a hard sphere, and the SPT shows that it always takes a positive energy to make a cavity. This positive energy then leads to our observed temperature dependence via (5.4). This seems about as far as we can take a simple model for explaining gas solubility. The usual recourse at this point is to introduce solubility parameters that are 158 characteristic of the solute-solvent combination. 33 ’ 84 Although this leads to very accuarte predictions over a wide range of temperature and solute concentration we shall leave it to the chemical engineers. 7. Empirical Analysis and Predictions Although analytic techniques give only limited understanding of these solubility data, as discussed in the previous section, ’empirical techniques may be applied usefully to them as we demonstrate now. The idea underlying our empirical analysis is to construct the thermodynamic quantities Aug, Ahz, and A3; for each solvent as the sum of contributions from the functional groups which make up the solvent molecules. For our 45 solvents we shall attempt to decompose Au; into Xe interactions with the functional groups: CHz (in linear molecules), C83, 08, C008, CEO, and CH2 (in cyclomolecules). Figure 30 is a plot of the excess chemical for 40 solvents verses number of C32 groups at 25°C. At first glance this does not look like a promising candidate for separation into functional group contributions. One might hope that the addition of a Cflz to any molecule might bring about the same change in excess chemical potential in each homologous series, but this is clearly not the case since the slopes for each series are so different. In fact for the polar molecules Au; turns upward as we approach zero 082 groups, while for the nonpolar alkanes and cycloalkanes the chemical potential becomes more negative (this holds for the 159 Apz', (cal/Incl) -400 —600 —800 -1000 160 0 Number of CH2 groups E I I -0 Alkanes : _ ’1 Alkanols 1 — Cl Acids _ — <> Cyclo — _ + Aldehydes ‘ 5' ..J 5 10 15 20 Figure 30. Excess chemical potential Au:=-RT£nL 2; number of C14z groups at 25°C. 161 perfluoroalkanes as well). In order to get some insight into what is happening we must look at our expression for * Apz; i.e. equation (6.35): 2 a ’ °° Aug = I; Io4npigaz(r,l)u12(r)r dr d1 . (6.35) As one increases the number of CH2 groups there are several parts of Eq. (6.35) that vary; the solvent number density , p1 (as illustrated on Table IV and Figure 23), the pair correlation function 812’ and the intermolecular potential u12(r). There is a great deal of evidence from computer simulations that the pair potential can be broken into group contributions. If we assume this is true and that the correlation function changes slowly, we may be able to appropriately scale our data. Figure 31 is a plot of the data, for 37 solvents,.fram which one may obtain group contributions for Aug. The ordinate is Aug/p1 at 25°C and the abcissa is the number of CH2 groups in the solvent molecule. The four sets of points on Figure 31 which give data for alkanes, alkanals, carboxylic acids, and alkanals, can be fitted to straight lines which have approximately the same slopes. This makes empirical analysis possible. One therefore takes the weighted average slope of these four lines to be the contribution to Aug/p1 of each CH2 group in a linear molecule. The analysis omits solvents which have one or zero CH2 groups, but the data for these are included 162 on Figure 31. This quantity, called 6“ (cal/mol)/(mol/liter), is given at the upper left of Table IX. In like manner the average slope of the curve for cycloalkanes on Figure 31 gives the corresponding 6 for an additional CH2 group in a cyclomolecule. The perfluoroalkanes’ data are not good enough to be analyzed this way because the samples were mixed isomers. If one extrapolates on Figure 31 the points for alkanes to the limit of zero CH2 groups one gets as intercept the contributions to Aug/p1 of the two CH3 groups; i.e. 2€P(CH3)’ at the ends of these molecules. The intercept on the ordinate axis obtained by extrapolation of the alkanol data is €H(CH3) + eu(OH), from which one may obtain £p(OH) by subtracting off eu(CH3) previously obtained. In similiar fashion we obtained 6u(COOH) and 6u(CHO). Values for all six of these en’s are shown in Table XI. The formula by which one reconstructs an estimated Au; at 25°C is, say, for an alkanol with, say, m CH2 groups: * . Ap2(est) - p1[ eu(CH3') + meu(CH2,lin) + eu(OH) ] , (7.1) from which L(25°C) may be obtained from Ap;(est)= -RT5n L. Figure 32 shows values for average enthalpy obtained from our solubility measurements. The ordinate is MEI/p1 and the abscissa is, as in Figure 31, the number of CH2 Au,” / ,01 (cal/mol)/(mol/liter) 163 O ‘ I I’ I F’I LI I I I I I I I I I I f O X alkanes 4 - I O O alkanals ‘ -50 — I Q I acids - I . . A alkanals 4 ' O cycloalkanes - Z 5 . e j -100 —- f o ' 0 —— .. x o 3 L. X .. - x .. .. X .. -150 — x x - _ .X _ x- I- x 4— )— X X -. I. X X "‘ "ZCN)‘_' ‘x "' .- l l L I l l L .1 l l l l l L l L I d 0 6 12 _ 18 Number of CH2 groups Figure 31. Experimental data for Xe in 37 solvents. Excess chemical potential divided by golveht number density y; number of CH2 groups at 25 C. A1730" / p, (cal/mol)/(mol/liter) 164 -100 _ ‘1 ' ' ' ' lz'. T I I r” * I I I I I - l . l I - . X alkanes : _200-_ 2 e alkanols .7 _ ' O O s acids —' .. . A alkanals : - . O O cycloalkanes_ 43007 X X x e ._- .. . -( .. x e - .. X a - -400— X _" -. x .- - x _ .. x : -500H— x —~ - >< : "6‘H3- I I I I I I I I I I 1, l I I I IX I 1 0 6 12 Number of CH2 groups Figure 32. Experimental data for Xe in 37 solvents. Excess enthalpy divided by solvent number density 11 number of CH: groups . ...-0 CD T115?" / p1 (cal/ mol)/ (mol/ liter) 165 0 L- I r I I I I I I I r I I I I I I 0 X alkanes - I 0 O alkanols .. -100 . . I acids —n " ‘ a A alkanals - - I O c cloalkanes- _ ' A . y 3 _. x x Q J -200 — Q __ - I ' I .. Q .. _ C .. -300 I:- X x —‘ .. X .. _ x >< 4 -400 I- l l l l l l L l l l l I l l l >l< l d 6 12 18 Number of CH; groups Figure 33. Experimental data for Xe in 37 solvents. * TAsz divided by solvent number density 21 number of CH2 groups. in 166 Table XI. Empirical Energy parameters for Xe solubility 37 organic solvents. Group Energy contribution (25°C) (cal/mol)[(mol[liter) ‘p ‘H ‘76 CH2 (lin) -7.45 -28.02 -20.36 CH3 -47.67 -99.01 -50.21 0“ +18.41 + 4.64 -14.89 COOH +10.16 - 9.57 -20.86 CHO + 7.22 + 3.02 .- 5.32 CHZ (cycl) - 2.98 -22.09 -19.11 solvents at 25.0°C. Table XII. Predicted igugxperimental values of Ostwald solubility for Xe in selected organic Predicted Experimental Solvent L (25°C) L (25°C) n-ClZHZ6 3.50 3.39 n-C7H1308 2.42 2.44 n-CachOOH 2.52 2.70 n-CAH9CHO 2.69 2.99 167 groups in the solvent molecule. For these quantities, as on Figure 31, the curve for alkanes, alkanals, carboylic acids, and. alkanals, are approximately' parallel straight lines. One obtains the enthalpic contribution eh for each of the six component functional groups by a similiar analysis as for 6“. The values of Ch obtained in this way are given in the middle column of Table XI. The formula by which one constructs AB: is, say for the same alkanol as in equation (7.1): Ah:(est) = p1[ eh(CH3) + meh(CH2,lin) + eh(OH) ] . (7.2) The A52 values predicted by equation (7.2) are averages over the experimental temperature intervals; nominally they correspond to about 25°C. Figure 33 shows values for the average entropy obtained from our solubility measurements. The ordinate is TA§:/p1. It has been multiplied by the average temperature T=298.15 K so that the ordinate’s dimensions will be (cal/mol)/(mol/liter) the same as on Figures 31 and 32. The componental entropy contributions 5T8 are shown on table V for the six functional groups. Values of 518 were obtained by applying to the data on Figure 33 the same technique as before. The formula by which one constructs A3: from these componential contributions is: A§:(est) = pal T296[€7§CH3) + meT§CH2,lin) + £1§OH)] , (7.3) 168 for the same solvent as in equations (7.1) and (7.2). Finally, one may estimate the temperature dependence of * Auz, over the temperature intervals near those for which the original data apply, from: It -* -* Apz(T,est.) = Ah2(est.) - TAsz(est.) . (7.4) In applying this equation, one uses equations (7.2) and (7.3) along with eh and 818 values in Table XI, and one obtains an estimate of solubility L(T) for Xe is whichever solvent that has been reconstructed. As a test of the ability of equations like (7.1) to predict Ostwald solubility, we show on Table XII a comparison. between. predicted. and, experimental values for five solvents, each chosen from about the middle of' its series. The predictions are obtained from L(25°C) =exp[-Au;(est/RT298)]. The predicted values of L(25°C) are within 12% of the experimental values for all of our solvents with two or more CH2 groups. The probable explanation for deviations at low carbon number is that the solvents are becoming more ordered and the correlation function begins to change rapidly at short chain length. The above test is largely one of consistency rather than predictability. But if suitable data were available, one could test whether our 6" values could be used to predict Xe solubility in, say, dicarboxylic acids, polyalcohols, and 169 other solvents . Finally, Figure 34 shows a plot of (Au; - 3mm)“;1 = shat/p1 verses number of CH2 groups. An extrapolation of the cycloalkane data to zero CI-i2 groups gives 81m: 0 , as one would expect. We note for completeness that the straight line fits are somewhat better than those in Figure 31. gm/ ,0, (cal/ mol) / (mol/ liter) 170 0 ‘7 I I I I I I I I I I I I I I I I ‘ X alkanes - —400 ‘ x . O alkanois . A x . a acids - ' a . x A alkanals - - O cycloalkanes - __ _. e A x ’ _. 800 _ I . x _ .. O x _ ' e x _ 4—1200'—‘ ~ O X -— - . x a .. . X _ — x -( - 1600 —- x “ _ x - x _ _2000 I I I I I l I I I I I I I I I I 0 6 12 18 Number of' CH2 groups Figure 34. giM/p‘, a number of CH2 groups for Xe in 37 solvents. 8. Conclusions In conclusion we discuss briefly how this work can be used to predict and understand solubility data and what further experimental and theoretical work would be useful towards these ends. The empirical results associated with Table XI and Figures 31.32.33 are suggestive but need testing on other systems before they can be- relied on for predictions. Some other C-, H-, and O-containing solvent functional groups for which the values of e", eh. and 61,8 would be interesting are, e.g. phenyl and its derivatives, carbonyl,alkoxy, etc. If solubility predictions from these prove to be robust one could extend the technique to find contributions of groups containing other elements, say halogens, N, 8, etc. A parallel. direction for empirical development is solubility of other inert gases and, beyond that, gases that are isaelectronic, or nearly so, to inert gases, etc. We anticipate that the en’s, en’s, and crs’s empirically found for different gases with respect to a single solvent functional group would be simply related to the Lennard-Jones potential parameters of the solutes.71 Finally there is some indication that the mixed experiaental-theoretical term (-RT£nL - g°")/ p1, which according to equation (6.54) equals gun/p , is meaningful i for espirical and analytic purposes. As more data become 171 172 available, its use for quantitative predictions can be explored. There are several interesting reasons 'why' analytical approaches to these data are difficult. In the following paragraphs I would like to summarize some of our observations on this topic. The quantities which currently lend themselves to theoretical analysis are the free energy of cavity formation, g , and its associated enthalpy, h , and CIV OIV entropy, s , and the energy of interaction u . cav . int Scaled particle theory provides a formal procedure by which one may calculate g?.v. There are several alternative theories which predict cavity free energies most notable of 85 This which is the surface tension theory of Sinanoglu. approach requires a calculaton of the surface energy associated with cavity formation in a solvent. The surface tension’s contribution is modified to take into account the microscopic cavity size. Recent reviews have shown that this theory tends to predict cavity energies which are too high, and is inconsistent with pure liquid heat of 72 vaporization data. In earlier papers we tested it on the perfluoralkanes and found cavity energies to be 2.5 times higher than the SPT prediction.65 The general consensus is that Sinanoglu’s theory can be used to get rough approximations of the cavity terms but SPT is much preferred when the parameters necessary to apply the later are known or can be evaluated.72 173 Boublik gt 9;. have developed a version of the SPT for nonspherical hard. molecules; such. as spherocylinders (cylinders with spherical endcaps) and other convex 86 At first this may seem like a better alternative bodies. to describing solubility in alkanes, unfortunatly, this theory cannot yet account for mixtures of particles of different shapes.14 More work in this direction may prove to be useful. Although scaled particle theory may be well suited to predict cavity energies there are restrictions to its application. The solute should be as hard and spherical as possible and must not have strong directional interactions with the solvent molecules. An example of an inappropriate (but important)solute would be a protein, for which a cavity would be an ill defined quantity. The solvent can be any real liquid. However,the effective solvent diameter,' a1, must be determined with great accuracy. For nonpolar solvents one can evaluate a1 from heat of vaporization data through equation (6.87). For polar solvents this equation is invalid and one usually has to use solubility data to evaluate a1.71 I The fact that the SPT works at all is actually quite amazing and. is ‘probably the result. of’ many' errors that cancel each other out. This means that we should be cautious when looking at higher order terms; i.e., the enthalpy and the entropy. We have demonstrated that the SPT does not predict the entropy of salvation. This is contrary 174 to previous thought. The reason this has gone unnoticed in the past is because most workers choose the mole fraction 55,74 scale. The chemical potential on the mole fraction scale is determined from [equation (8) from ref.55]: * Apz = -RT5nx2 Auz + RT6n(RT/v1) , (8.1) where v1 is the molar volume of the solvent. To find the corresponding entropy we take the derivative with respect to temperature: -# As2 = As2 - R + RTcIc - ROn(RT/v1) . (8.2) The extra terms in equation (8.2) are usually much larger than A5: but they have no physcial interpretation. For example, for Xe in hexane we have A§:= -4.4 cal/mol X, but A32: -15.8 cal/moi K. This excess pseudo-entropy serves to mask the deficiencies in sc ; i.e., instead of predicting IV As = 3391': 0.29 cal/mol K, we would say As: s8". spun“ = -ll.1 cal/moi K, which doesn’t look so bad when compared to Asa. This does not necessarily mean that the SPT is wrong; it is possible that the rest of the entropy could be in the interaction term, g . int To settle these many questions one must turn to computer simulations. Using equation (6.36) we obtain : 1 mum. = (0(5) dg . (8.3) O 175 Swope and Andersen41 have used molecular dynamics to evaluate (8.3) for inert gases in water. The idea is to run your simulation and gradually couple a solute particle into the solvent. This is done by turning on the solute-solvent potential in a stepwise manner as suggested by Eq. (8.3) and calculating the coupling energy, U(g), at each value of 5 until the particle is fully coupled. One can then integrate the smoothed results to obtain the excess chemical potential. Once the particle is fully coupled it is possible to calculate the entropic contribution to the free energy by simply subtracting off the energetic part; i.e. a ‘ -:I: Apz = I 0(5) 65 m U(g=1) - 7432 . (8.4) o Jorgensenl has developed interaction site potentials which predict the properties of pure alkanes and pure alkanals. These are potentials in which the molecule is represented by a set of discreet interaction sites that are commonly, but not invariably, located at the sites of the atomic nuclei. For the alkane potentials Jorgensen has used the various functional groups as the interaction sites. These have also been used to predict the solubility of various alkanes and alkanols in water.87 While this data is certainly valuable one must note that solubility in water is a much more complicated process than in nonpolar solvents due to the complex structure of liquid water. In principle 176 it is possible to do the same calculation for inert gases in alkanes and alkanols and maybe other solvents. By' carefully choosing the way the solute - solvent potential is coupled it may be possible to gain insight into how repulsive and attractive interactions affect the salvation entropy. This in turn would clarify the deficiencies in analytic models and help us to correct them. To complete our experiment it would be interesting to look at Xe solubility in the lighter alkanes, i.e.; ethane, propane, butane and methane (despite the fact that methane’s boiling point is lower than xenon’s at 1 atm. The trick is that by using a tracer the solute gas is so dilute that it will not condense until a significantly lower temperature77 ). These are simpler solvents and thus may help bridge the gap in our understanding of solubility. Also, our experimental technique offers a unique advantage over other cryogenic experiments in that we can easily measure 89’90 This solubility in the limit of infinite dilution. makes the subsequent analysis much easier since one can neglect solute-solute interactions. Our technique may also prove to be useful in analyzing solubility near the solvent’s critical point (we again have an advantage over other methods since we can ignore solvent vapor pressure effects). A great deal of recent work has been along the lines of gas solubility near the critical temperature of watergl; performing complementary experiments on non-polar and polar systems such as ours could prove to 177 be very enlightening. 9.Introduction: Solubility of Nonpolar Gases in Water at Elevated Pressure. This section of my thesis reports solubility measurements of the. gases nitrogen, argon, krypton, and xenon in water. For each of these gases we measured, at 25.0°C, the dependence of gas solubility on partial pressure of the solute. The pressure ranges studied were approximately: 44-116 atm for N2, 22-101 atm for Ar, 33-81 atm for Kr, and 5-48 atm for Xe. The data are analyzed in terms of molecular theories based on thermodynamics, and statistical mechanics. These experiments are an extension of previous work on Xe solubility in organic liquids‘ and aqueous solutions.”’92 The study of gas solubility in liquids is an. old and well developed subject in chemistry and physics. For dilute solutions at sufficiently low pressure, the mole fraction 93 solubility, x2, is well described by Henry’s Law : P = k x (9.1) where P2 is the partial pressure of the solute gas and ha is a constant of proportionality. The so called Henry’s constant, k”, is characteristic of the solute-solvent system 178 179 at a given temperature. At higher pressures Eq.(9.1) must be modified to account for nonideality of the vapor. Lewis94 introduced a new form of Henry’s Law in which the solute partial pressure is replaced by its corresponding fugacity; f2=kflx2. At still higher pressures the measured solubility fails to obey this relation as well. It shall be shown that by measuring deviations from Henry’s Law one can obtain information on the salvation process. The importance of understanding the effects of pressure on gas solubility comes from its many applications. At an industrial level, extraction and separation processes are often performed at elevated pressure. In biology and medicine, interactions of inert gases with living organisms are also important. For example, the gas mixtures breathed at high pressure in deep sea diving consist mainly of inert gases. Interactions between these gases and the diver are responsible for decompression sickness and inert gas narcosis.10’95 Solubility limited phenomena are also relevant in marine biology. Several species of deep water fish have swim bladders in which partial pressures of nitrogen up to 10 atmospheres have been observed.96 In order to determine the mechanisms used by the fish to inflate its swim bladder, it is important to know the concentrations of dissolved gases available at a given depth. Such variation is generally determined by the hydrostatic pressure. Another motivation for this research is to look at the 180 role of hydrophobic effects on gas solubility. Hydrophobic interactions are important for understanding the stability of biologically important macromolecular structures. Hydrophobic effects are thought to play a key role in 97,98 protein conformation and enzyme specificity. Several experimental studies have investigated the behavior of simple polar and nonpolar organic solutes in water.99’100 Inert gases would in some sense be the prototypical hydrophobic solutes because their interactions with water are very weak compared to those of water with itself. The inert gas elements have been studied extensively as 77,78 prototype solids, liquids, and gases, so many of their interactions are well known. The low pressure solubility of these gases in water have also been determined 6’32 With the advent of improved analytic73’101 41,102 accurately. and computational techniques it has become possible to predict these solubilities with good accuracy. By providing data in different ranges of temperature and pressure one can test the robustness of these theoretical predictions and make appropriate refinements. 10. Experimental The two most common units of gas solubility are the Ostwald (L) and mole-fraction (x2) solubilities. They are defined as: I.) 92 n2 3 fl and X2 3 W e (10 a 1) pz 1 Z Ostwald solubility is the ratio, at equilibrium, of the concentration of dissolved gas molecules in the liquid solvent to their concentration in the gas phase. The mole -fraction solubility is the equilibrium fraction of solute molecules in the solvent. In recent literature measurements in terms of the Ostwald solubility are characterized as being on the ’number density-scale’.14 Figure 35shows a schematic of the experimental apparatus. The Figure is divided into two parts to show each step in the measurement process. First the liquid is saturated with the solute gas at high pressure as depicted in 35a. Once equilibrium is attained the liquid is isolated and the sample cell is transferred to the analysis system, 35b, where the amount of dissolved solute in the pressurized liquid is determined. Together, this 181 182 To Vacuum Pump v...mooooooooooooooooooom (b) Analyzer 1 Pressure Transducer [lflll hmm 3 Figure 35. Schematic of experimental setup (a) high pressure equilibration system (b) analysis system. 183 constitutes a modified Van Slyke method 103 in which the sample volume is the entire liquid volume. Our technique is different from Van Slyke’s in that it allows us to measure directly the mole -fraction solubility, Ostwald solubility, and partial molar volume. We briefly outline some experimental details below. The sample cell is a 100cm3 stainless steel volume with a ball valve to isolate the liquid and allow for easy transfer between the equilibration and analysis stages of the experiment. The solute gases we used were obtained commercially. Their purities were: N2(99.99%), Ar(99.998%), Kr(99.997%), and Xe(99.996%), (N2, Kr, Xe from Linde Gas Products, Ar from Matheson Gas Products). At the beginning of the experiment we load the sample cell with twice distilled water which has been degassed using standard methods. The water is loaded under vacuum to prevent air packets from forming. The cell is weighed and the mass of the water, after making proper .buoyancy corrections for the weight of displaced air, is denoted mo. To help prevent gas spaces from developing in the cell upon liquid compression we add some extra water above the ball valve. The mass of the additional water is Am+. The cell is then connected to gas volume V8 and pressurized. Mixing is achieved with a specially designed magnetic stirrer. Gas pressure is monitored with an accurate capacitive transducer (Setra Systems), and. a pressure gauge (Heise Inc.) for calibration. The sample cell and gas volume are submerged 184 throughout the entire experiment in a constant temperature water bath at 25.0:0.1°c. After some time, typically 2-3 days depending on stirring speed and solubility, the pressure readings reach a steady state. To be sure that equilibrium is attained we monitor the pressure for an extra 8-12 hours. At this point we record the final pressure of the solute gas, Peq’ and the ball valve on the sample cell is closed. The gas volume is then slowly depressurized. Excess liquid above the ball valve is collected and weighed. This mass is denoted Am-. The water remaining in the sample cell is now: In = m + Am - Am . (10.3) The analysis system as illustrated in Fig. 35b consists of a detachable gas volume, V8a , with a pressure transducer and mercury manometer. The gas volume is variable in the sense that different sized cells can be attached to the side arm of the analyzer. We choose the gas volume such that the pressure of the evolved solute will be within the optimum range of our sensing devices. This pressure can be estimated. via Henry’s Law with. well known low pressure 6’32 For the solute gases and pressure solubility data. ranges in this experiment we used two analyzer volumes, namely: slBOcc (N2 and Xe) and s360cc(Ar, Kr, and Xe). To measure the amount of solute in, the saturated liquid we connect the sample cell to the analyzer and evacuate the 185 upper volume. The main valve is then opened thus releasing the dissolved gas into V8a shown in Fig. 35b. With stirring a steady value of pressure is usually reached within 2-6 hours. The final pressure of the solute gas in the analyzer, Pa’ is found after correcting for the vapor pressure of the water.79 To calculate the mole-fraction and Ostwald solubilities we first must find the density of the solute gas, 928(P), at Pa and Peq' Accurate virial expansions were used in the appropriate pressure ranges for N2 , 104 Ar , 105 Kr , 106 and Xe . 107 One can then find the number of moles of gaseous solute in the analyzer to be: - 8 a (2) n2 - p2 (Pa) [Vg + V8 ] , (10.4) where V;Z)is the extra. gas volume that arises from the decompression and outgassing of the pressurized liquid. Assuming the low pressure liquid density is that of pure water we have: (2)_ _' a v8 "V6.14 In"20 /p"20(25.0 C) , (10.5) where V is the volume of the sample cell and m is cell H20 defined in Eq.(10.3). To find the total number of moles of solute we must also account for any remaining in the liquid. Adding this correction we find for the total amount of solute in the sample cell at Peq 186 n2,total= 928(P8)[ [vga+ V;2)] + Lo(mlIZO/p"2°) ] , (10.6) where L° is the low pressure limit of the Ostwald solubility 6,32 at 25.0°C. The number of moles of water in the sample cell, n1, is the mass of the water divided by the molecular weight, i.e., n1 = IIIn 0(g)/18.01534(g/mol). One can then 2 easily calculate the mole fraction solubility from Eq.(10.2). At the equilibrium pressure Peq the liquid volume is fixed to be that of the sample cell, therefore the number density of solute molecules in the liquid is p21: Dividing this quantity by the gas number n / V . 2,total cell density at peq one obtains the Ostwald solubility, L, as given in Eq.(10.2). One may also use this technique to obtain a rough estimate of the solute molar volume by: 325: [Vc.”- [mazo/pnzo(25 C°,1'atm)]exp(-x(P.q- Pa)]/ n2, (10.7) where x is the isothermal compressibility of pure water,108 x = 4.571310“5 atm-1, and n2 is given by Eq.(4). Essentially this expression means that if we take Va”, the known volume of the pressurized liquid, and subtract off a term due to liquid compression then the remaining volume, which is in the square brackets of Eq.(10.7) is due to the solute molar expansion. Dividing this by the amount of solute in the liquid gives the molar expansion produced per mole of 187 solute added, i.e., 62. Since our solutions are very dilute, typically xzarIO-‘, we assume in this development that the density is the same as for pure water. This assumption is also supported by the fact that our results agree fairly well with literature values.109’110 We stress that the strength of this technique is that by fixing both the sample volume and mass, one may easily find the Ostwald solubility and the molar volume. Other methods usually sample only a known mass of saturated liquid and therefore are limited in that they measure only mole fraction directly. Separate measurements are then needed to calculate the solubility on the number density (Ostwald) scale. It has been argued that the number density scale has certain advantages over the mole fraction scale in analyzing 99 The results from a statistical mechanics perspective. primary weakness of our current technique is that the measurement takes a relatively long time. A better alternative wouLd be to use small samples of fixed volume 111,103 This from a large reservoir of equilibrated liquid. is a more traditional approach with the important addition that we know the sample volume at Peq. 11. Results The experimental results we obtained are shown in Table XIII. The first column gives the pressure at which the measurement was made, the second column gives the corresponding fugacity. The third and fourth columns tabulate solubilities on the mole fraction and number density (Ostwald) scale respectively. We also include the average value obtained for partial molar ‘volume of the solute. The fifth and sixth columns give the data in terms of a new solubility parameter which will be discussed below. Judging from reproducibility and by comparison with other datas’112 a conservative estimate of our fractional uncertainty in solubility is about 1-1.5%. The primary sources of error are pressure measurement(:2.0 psi for high pressure, $0.5 Torr for low pressure), volume measurement (t0.l% reproducibility), and microbubble formation in the pressurized liquid. The first two terms lead to an uncertainty of less than 0.2% so it seems that the last term is the main contributor. Fugacitiea were calculated from volumetric properties of the real gas via20 P f = p exp[ 'fi; I [v(p’) - 3%? ldp’ ] . (11.1) 188 189 The integral in Eq. (11.1) measures the cumulative deviation of the real gas volume, v(p’), from that of the corresponding ideal gas at pressure p’. To obtain v(p’) we used the virial expansions referred to in the previous 104-107 section. The fugacity is related to the chemical potential per mole of the gas by 126: 12 A3 p29: Run[ J , (11.2) k T where f2 is in atmospheres and k=R/NA= cma-atm/molecule-°K and A=h/(2nka)1/2is the thermal wavelength of the solute.16 Figure 36 shows a plot of the equilibrium fugacity versus mole-fraction solubility for our data. Also included is data for N2 in water at 25.0°C from Ref.(112). As one can see, the agreement with our measurements is good. The curves in Fig{ 36 have the common feature that the mole—fraction solubility increases approximately linearly with fugacity. Table XIII shows that the Ostwald solubility tends to decrease with f2. If the system was in complete accord with Henry’s Law we would expect the quantity ku=flx2 to be constant for each gas. Figures 37(a-d) show such plots with respect to solute concentration. For N2, Ar, and Kr there is a clear increase in k" by as much as 7-10% in the concentration range studied. For xenon there appears to be very little variation in kn' A problem that we encountered with xenon was that above a pressure of 17 190 atmospheres the solubilty decreased rapidly and the molar volume increased to about 125cm3/mole. The data point from Table XIII at 23.33 atm seems to be anomalous in the sense that the solubility has decreased sharply ,compared with lower pressures, but we did not see a large molar expansion. We associate this effect with the onset of clathrate formation.113 This speculation was confirmed in separate experiments with xenon-water mixtures at Izq>30 atm. Instead of subjecting the equilibrated sample to analysis we quickly depressurized the cell and pour the liquid into a beakeru Several ice-like crystals, as large as 5mm in diameter, were observed. These melted away within minutes. We have included on Table XIII the solubility measured during the clathrate phase but we are not sure that the results are reliable since it is known that this is a glassy state which may require very long equilibration times.113 It might be worthwhile to repeat our low pressure xenon measurements because the pressure ranges we explored were at the lower end of our sensing accuracy. Our technique using radioactive tracers which would be useful toward this end.18 191 Table XIII. Experimental results for splubility of N2, Ar, Kr, and Xe in water at 25.0 C. The second column is the solute partial pressure. The third column is the solute fugacity. The fourth and fifth columns are the mole-fraction and Ostwald - solubilities respectively. The sixth column is the new solubility parameter r. The seventh column is the excess chemical potential of the solute in the liquid. The eighth column is the experimental average value of partial molar volume. Solute P(atm) ftiatm) x'I-lo‘) I. r 781524151) V'ica'laoi) 1‘: 6' 44.55 44.25 4.902 0.01465’ 0.01501 2466 g 50.46 50.10 5.462 0.01466 0.01464 2495 , 63.66 63.17 6.609 0.01446 0.01463 2503 77.26 76.63 6.100 0.01423 0.014355 2514 31 z 2 ‘ ’ 93.30 92.54 9.562 0.01397 0.014045 2527 ‘“‘ 102.19 101.36 10.32 0.01360 0.01363 2536 115.60 114.99 11.46 0.01360 0.01356 2546 Ar 22.505 22.190 5 426 0.0322 0.0331 2019 25.50 25.01 6.136 0.0321 0.0331 2019 36.12 37.24 6.993 0.0313 0.0327 2026 44.36 43.20 10.36 0.0309 0.0325 2030 55.95 54.11 12.61 0.0301 0.0321 2036 ,3 2 1 66.66 66.15 15.41 0.0292 0.0316 2047 71.67 66.73 16.06 0.0293 0.0317 2044 79.19 75.65 17.43 0.0267 0.0313 2053 67.90 63.61 19.06 0.0262 0.0309 2059 101.27 95.70 21.59 0.0276 0.0306 2065 x: 33.30 31.01 13.45 0.0506 0.0566 1679 40.64 37.26 15.91 0.0465 0.0579 1666 41.16 37.69 16.22 0.0461 0.0563 1663 52.14 46.64 20.03 0.0463 0.0562 1665 3° 3 2 56.66 51.93 21.53 0.04345 0.05625 1705 69.73 60.06 25.13 0.0417 0.0567 1700 60.56 67.91 27.67 0.0391 0.0556 1712 60.66 67.99 27.79 0.0369 0.0554 1714 x0 4.791 4.669 3.592 0.0994 0.1041 1340 6.344 6.103 4.716 0.0962 0.1046 1336 6.151 7.600 5.979 0.0950 0.1036 1342 9.415 6.947 6.930 0.0950 0.1046 1336 ,7 g 2 11.25 10.56 6.069 0.0917 0.1034 1345 12.10 11.33 6.750 0.0915 0.1045 1336 14.69 13.55 10.36 0.0660 0.1034 1345 15.74 14.44 11.035 0 0669 0 1034 1345 --_- x. 23.33 20.46 13.60 0.066 0.090 1426 44 32.76 27.19 17.46 0.057 0 0664 1451 117 33.29 27.52 17.62 0.057 0.0659 1454 125 47.69 35.99 21.17 0.040 0.079 1506 131 192 120 ' i ' i I T V I l I r 7 —. '_' o 4 - bfitrogen « 100 - 9 _. .- 0 .1 _ o . so ' ° : ,\ - O o Argon _: g ' 0° :7 ~ d ' O 4 v 60 — a -— r? ‘ 9 o ' C ‘ 3 :_ O o a 7 go 4:0 _ 0 cm Krypton 1 2. : a 7 - 1? 00 I 20 :— ‘(enon 1 _ xx XX ‘ . .J .. xx . x x O t. ' l ' l L 1 ' ' l ' ' I ' q 0 0.001 0.002 0.003 Mole Fraction. x2 Figure 36. Solutg fugacity u mole-fraction solubility at 25 C. 193 :2. .— “:2 2 N2: 3 I .44.|194 l4!-44|49-ll‘l¢4’ incl — 4 6 I... ... :23haz 3 IL L 3 Q A 3 I I e u 3 ._ 3 4 3 U 3 I .- ..L L — n p p tr — u h an ...3._us...m v.32 n3: 3 N33 3 .33 3 fl 7 q . . - iJ - )3 A 8, 1 a ) 1 do”: 2 L l 5 r I o L 1 o o A o A $3M” 3:03“... «5353 aaonu .5ch seem ocean 3333 v 33:. 325? .52.. v «56.. v scam v (um) ‘V‘MX (um) “it/'12“! a _ a: : 0:33 3 .33 3 3 |.oJ-44ll‘fll'—. .4 1'4 0|" .d. . 1.]?!4']. :38”— I .. 1 Sam. 4 L v n.0an.” L - 4 I. scan. r ‘ a . . . m l < 1 :92”- 1|.— . . .-- _1-rh L . _ a . . :aNfl— ...x ...ozusi Box «.363: :53. @0333 c - i _ - 1 . . a .J . .5600 1 m J 3 30:93.2 . o 1 250m .. a O L 1 a 1 88.: L L 4. r ..I..3.Iplr.— - b p b Fill-I. rlrlll gna- (um) 'V‘h'x (um) “cm-"74 s constant 2:, mole-fraction I °c. Figure 37. Henry solubility at 25 12. Pressure and Concentration Dependence Deviations in Henry’s Law, such as those shown by our data, are best analyzed in terms of the solute chemical potential. In our experiments the vapor pressure of the solvent, P(azo,zs°c) a 23.76 torr, was small compared to the solute gas pressures and can be ignored. The chemical potential per mole of gaseous solute can then be written as: O fz(T,P) 2 2 p¢ where 142’”) is _the standard chemical potential of the gas at temperature T and pressure P’. The choice of reference pressure is such that the gas behave ideally at P’. We shall follow the usual convention and set P0 equal to 1 20,22 atm. This expression for 142' is identical to Eq.(12.l) but will prove to be' easier to work with in subsequent discussions. The chemical potential of the solute in the liquid, "21’ is completely described by the variables T,P, and x2. At constant 'I‘ one may express the differential change in ”2‘ as 114 194 - auz d auz - x2 + dP . (12.2) PT 36 ‘I' To understand how "21 depends on solute concentration we use the Kirkwood-Buff (KB) solution theory.115 This is an exact formal treatment in which the solution properties are obtained from integrals over 'pair~ correlation functions. For two components the KB integral is defined as: m G = I [g (r) - 1]4ur2dr , (12.3) ii 0 ii where g‘j(r) is the orientationally averaged pair correlation function between species i and j, and depends only on intermolecular separation r. To relate these quantities to pZ'CT,P,x2) we use Eq. (22) of Ref. (115): 1 1 [ a "2 J = 1 + 91(2612 Ga: 622) , R T 0 x2 P,T x2 1 + p1x2(611+ G22- 2G12) (12.4) where 91 is the molar concentration of the solvent. For dilute solutions Eq. (12.4) can be expressed in terms of the infinite dilution limits of the molar volumes and solvent isothermal compressibility, x1 26: 196 l 6 p i 1 1 __ [ 2 J = —— + O [ -G + V O - 2V 0+ RTX O] R T a x 9,7 x v 22 1 2 1 2 2 1 + 0(x2)2 . (12.5) Throughout this thesis we use the superscript o to refer to the infinite dilution limit. Integrating Eq. (12.5) and keeping terms to first order in x2 yields 116: 1 _ o 2 112 (x2.T.P) - 112 (T.P) 7‘ RTlmxz + XZE(T) + 0(x2) . (12.6) where O _ _ _ o g(T)-RT( 622+ v1 2v2 + RTx1°)/v1° , (12.661) and p20(T,P) is a constant of integration. The KB integral. 622, plays an important role in determining the concentration dependence of the chemical potential. It is a measure of correlations among solute particles in solution. This can be illustrated by multiplying 622 by the solute concentration pz: 0 _ _ 2 p2622 - Io pz[ g22(r) 1 ] 4nr dr . (12.7) This quantity reflects the total average excess (or deficiency) of solute molecules surrounding a fixed solute 197 with respect to a uniform distribution.99 For example, if 622 is positive the solute molecules are more correlated than if they were distributed randomly. This sort of solvent induced clustering in aqueous solutions is commonly referred to as a hydrophobic interaction.97 An exaggerated example of such a situation would be micelle formation in detergent solutions.117 A driving force in this example is the fact that the detergent has both polar and nonpolar regions which compete to minimize the free energy. The resulting conformation is that solute molecules cluster together with polar regions in contact with the water and the nonpolar regions in contact with each other. This sort of mechanism plays a vital role in biological process such 118 It is as membrane formation and pmotein conformation. not obvious that such mechanisms should come in to play for nonpolar gases, particularly those we have studied, since these solutes have little structure, i.e., the solutes are mono- or di- atomic and cannot be divided into polar and nonpolar regions. It has been shown however that solvent induced interactions can lead to clustering in hard sphere mixtures where there are no attractive interactions at an.119 A more familiar occurence of 622 is through its relation to the osmotic pressure of a solution. By analogy to the virial expansion for real gases one can show that the osmotic pressure ([1) of a two component solution obeys an equation of the form 22: 198 n/RT = p2 + B p + B p + ... . (12.8) N* The second osmotic virial coefficient of the dissolved solute is 126: 32 = -TG . (1249) * The value of 82 is frequently used in statistical theories of polymer configurations as a measure of the interactions between segments along the polymer chain.120 The pressure dependence of the chemical potential is described by the well known thermodynamic relation 20: 6112l [__) . .2 , (.2...) BP 71 ‘1' i! where v2 is the molar volume of the solute in the solvent. Applying this to Eq. (12.6) allows us to evaluate pzeas: P , )1 =1 v (p') dP’ + C(T) , (12.11) 0 2 where C is an integration constant which depends only on temperature. If the solution is well below the critical temperature of the solvent, the molar volume is weakly dependent on pressure and can be taken outside the 199 integral.121 We then substitute its value at infinite dilution, v20. Putting this result in Eq. (12.6) gives: 0 - 2 p - C(T) + v2 p + 31‘0an + ngvr) + 0(x2) . (12.12) The equilibrium condition between the vapor and the liquid is “29 = 02'. Using Eq.(12.l) for p29 and Eq.(12.12) for pal gives: f (T9?) 02°(T) + R'wn[ z—— ] x P¢ 2 = C(T) + P v20 + ngw) + 0(x2)2. (12.13) In the limit of low pressure and infinite dilution the bracketed term on the left hand side of Eq.(12.13) reduces to Henry’s constant kuo, so that one may evaluate C(T) - 15¢(T) = RTbn(k"°). This leads to the desired result: f2(T,P) ] o x P¢ 2 Run[ = Monwf) + Pvz + ngvr) + 0(x22) . (12.14) This expression was previously derived in Ref. (116), but we have filled in some of the steps. The observed deviations from a constant kno in (f/xz) can be attributed to increased Pv work in the liquid, and possible solute-solute 200 interactions. In the limit x2§(T)-40 we obtain a well known result in chemical engineering: 00(1‘ /x P") =bn(k °) + 2 (12.15) 2 2 H RT Equation (12.15) is known as the Krichevsky-Kasarnovsky 112’121 For the gas-liquid systems in this paper equation. (Pv‘o) is typically 20f100 times greater than (szT). Therefore, by Eq. (12.14), we will only be sensitive to g(T) if G22 is significantly larger than the molar volumes, i.e., g(T) a: XZRT ( C - Gzzlvao)’ where C is of order (-3) for typical solute molar volumes and solvent compressibilities. 13. Data Analysis Equation (12.14) can be used to obtain information on the pressure and concentration dependence of gas solubility. Figure 38 shows a plot of the quantity [cnlfz/sz¢) - PvzolRT] m x for each of our solutes. A linear 2 regression fit to the data gives for the ordinate axis intercept the value of mug) and the slope determines E(T)/RT. From g(r) and Eq. (12.6a) one may then find 622. The calculated slope was very sensitive to the choice of molar volume so that reliable values of Ge: are difficult to determine. To illustrate this point, Table XIV lists values of 622 and 6n(ku°) obtained for each solute at for different assuemd values of v 2°. The values of v 2° are taken from Ref. 109, in which standard literature values are compiled, along with the authors’ own direct measurements. The values of Véo given in Table XIV represent the range of generally acceptable data. A common way to obtain the partial molar volume of gases in liquids is to use solubility data in .conJunction with Eq.(12.15). This obviously is not appropriate to our analysis, which relies on Eq.(12.14), since the resulting molar volume would give zero for £(T). o When possible we have used values of v2 which were determined directly. and have indicated when otherwise. For 201 202 xenon the only direct measurement of molar volume is by Biggerstaff gt __1,. at elevated pressure.110 The data on Figure 38 were fitted using the following intermediate 122 values of molar volume; 32°(N2) = 34.2cm3/mol, 32°(Ar) 32cm3/mol, 32°(Kr) = 32cm3/mol, and 32°(Xe, P<17atm) 43.50m3/mol (the average nitrogen value corresponds to a measured value in Ref 122). The corresponding values of respectively. On Figure 38 we have shown on the ordinate axis using filled symbols the experimental values of (11(kno) for each solute at 25.0°C as given in reference 6, viz., kfl°(N2)= 85251, kH°(Ar)=39746, ka°(Kr) = 22252, kn°(Xe) = 12885. The intercepts of our graphs are much less dependent on choice of molar volume than the slopes and agree with the known results very well. Thus, the values of k"0 that one obtains from our plot are: k"°(N2) 84669, k"°(Ar) = 39875 , 12925. We also found that k"°(Kr) = 22261, and ka°(Xe) the xenon data at higher pressures could be fit to the same curve as the pre-clathrate data, provided we used our experimental value of v2°(Xe,P>20atm) a 125cm3. It is interesting to note what happens when Va0 is treated as a free parameter which can be varied to obtain the best straight line fits by chi squared minimization. Ina/x...) — Pvz/RT 203 12.0 _ ‘ T T T ' ' i I I I 1 r 115 '— i 89—6-6—99-9 Nitrogen 11.0 % —9-0——-0——-Oo+—o—o 10.5 Argon 10.0 _ 3 $——o—°——a—a L. i~ Krypton 9-5 W )- .. . : Xenon ' 9.0 L ; L ' l L l L i l i L l 0 0.001 0.00 0.003 Mole Fraction. x2 Figure.36. dnifz/xz) - PVéo/RT x§,mole-fraction solubility at 25°C. 204 Table XIV. Results of Kirkwood-Buff analysis. The third column gives the partial molar volumeaof the solute used in Eq.(12.14). The fourth and fifth columns give the results for the KB integral Ga: and the log of ghe infinite dilution limit of Henry's constant, £n(k" ). Solute T(°C) vaoicm’/mol) 621(cm3/mol) (u(kaoi N. 25.0 33‘ —236 11.346 35.7‘ -27 11.347 hr 25.0 32"' +169 10.5935 Kr 25.0 33‘ +24 10.010 31' +69 10.011 x: 25.0 41‘ +196 9.4666 46’ +240 9.4669 He 0.0 15.5 -157 11.791 23.7 +927 11.601 29.7 +1720 11.609 He 25.0 ~ 15.5 -22 11.674 23.7 +996 11.661 29.7 +1745 11.666 N: 25.0 33- -4.5 11.360 35.7 +345 11.369 143 50.0 33 -92 11.604 35.7 +330 11.611 Hz 0.0 20 -16.5 10.953 25.2 +262 10.956 26.7 +343 10.960 6: 25.0 20 +7.0 11.170 25.2 +314 11.173 26.7 +403 11.174 11' 50.0 20 +3.0 11.242 25.2 +305 11.245 26.7 +392 11.245 ca‘ 25.0 34.5 -13 10.566 37.4 +223 10.596 036‘ 37.6 51 +125 10.595 53 +470 10.609 1h 205 In every case the optimum fit leads to |GZZ|<50cm3/mole. However the resulting value of the ’effective molar volume’, vz'”, is often not in agreement with experiment, i.e., we obtain by this procedure v2“! ( N2 ) =36¢malmol , vz'”(Ar)=26cm3/mol , v2°"(Kr)=28cm3/mol, andvz'”(Xe)=10cm3/mol. The value for Xe is certainly unphysical since Xe is known to have the largest molar volume of the four gases studied. Table XIV also shows the results obtained by application of this analysis to data of other worker on high pressure gas solubility in water. These data appear below the double horizontal line on Table XIV. The systems include nu25-100061m;0°c,25°o),123 “2(25-10008tl; 0°o,25°o,50°0) 112 112 124 , N2(25-10003tm; 25°C,50°C), c114 (0-6000p81; 25°C), 125 Czi-i‘5 (0-10,000psi; 37.8°C). For H2 and N2(50°C) the gas phase fugacities used were those of Demming and Shupe.127 For the remaining gases the fugacities were calculated in the manner described earlier; the equation of state data we used to calculate f2 is taken from references 128, 104, 129, and 130 for He, N2, 011‘, and C2116, respectively. Although ethane isn’t a spherical solute we include it because of its similarity to Xe in aqueous solutions; i.e. both gases have comparable solubilities at 1 atm .and both form aqueous clathrates at 25.0°c.6’ 1139 131 The analysis uses all the published solubility data except the first point of the methane data (P=34lpsia) for which x2 seemed anomalously high. Once again we see that the calculated values of 622 206 are strongly dependent on the choice of molar volume. Molar volume data are taken from Ref.(109). For He there is no direct measurement of molar volume. Of the three quoted values shown in Table XIV, two are obtained from solubility 0 132,96 measurements (v2 = 15 . 5cm3/mol , v2°=29 . 7cm3/mol) and the other is an estimate from the pure liquid triple point density.116 These data, also, have the general feature that one can choose an effective molar volume which leads to the best fit of the data. The effective molar volumes that one obtains are: vz'”(i-Ie)=l5.5cm3/mol, v2°"(H2)= 20cm3/mol, V2.tt ( N2) = 33cm3/mol , v2." ( CH4) = 34cm3/mol , and vz'"(C2H6)=50cm3/mol. All are close to the experimental results with the exception of He and H2. Watanabe and Andersen116 have performed molecular dynamics simulations of the Kr-water system. Their results indicated that the Kr solutes tended to undergo ’hydrophobic repulsion’ in solution. By integrating the solute-solute correlation function, as evaluated from ,computer simulations,they found. that. Gé2(Kr) = -1004A3/molecule = -604cm3/mol. This means the solutes tended to avoid one another in solution. In the appendix of that paper they include an analysis of solubility measurements which is similar to ours. They find a definite correlation between maxi) for a given solute and its corresponding second osmotic virial coefficient, B: of Eq. (12.9). The general tendency was that as max) increased, B: decreased. By extrapolating experimental results they postulated that 207 622(Xe)a -7750m3/mol, 622(Kr)u -362cm3/mol, 622(Ar)a +16cm3/mol, and G22(Ne)u +821cm3/mol. They assert that the less soluble gases (large kuo) are less polarizable and tend to demonstrate hydrophobic clustering. The more polarizable gases such as Kr and Xe are relatively hydrophilic and prefer to be surrounded by water. We see no systematic variation of this type. We believe that the systematic variation observed in reference 116 for He, N2, H2 and CH4 is attributable to choice of molar volume, and selection of data (for example, they neglect all nitrogen data above x2=0.003 at 25°C and above x2=0.004 at 50°C). Our results do not rule out such behavior, but within the experimental uncertainty of v2O we cannot support such findings. It is necessary to point out that the Kr solutions of Watanabe and 116 computer simulation were 5‘10 times more Andersen’s concentrated than the solutions we made. It is possible that at this higher concentration one starts to see clathrate behavior for krypton such as ‘we observed for 133 which form around the solute xenon. The ice-like cages in the clathrate phase could provide a mechanism for ’hydrophobic repulsion’ between solutes. Our results are more consistent with the theoretical calculations of Pratt and Chandler.101 They developed a theory of the hydrophobic effect which is based on an integral equation for the relevant pair correlation functions of the aqueous solution. The theory is more accurate than previous analytic attempts because it utilizes 208 the experimentally determined oxygen-oxygen correlation function of pure water as an input parameter. Table V of Ref. 101 tabulates the calculated values of second osmotic virial coefficient, B:, for hard spheres in water in and a hard sphere solvent. Although Pratt and Chandler’s model predicted some hydrophobic clustering of the solutes, the resulting values of 622 only ranged from -15 to 96 (cmalmol). SinCe a hard sphere is in some sense the limiting case of an inert gas solute with zero polarizability it might be reasonable to expect that 6:; would be the largest possible value for the inert gases. Unfortunately no attempt was made by the authors to calculate G22 for more polarizable gaseous solutes. To briefly summarize, we found that the solubility data could be accurately fit by taking account only pressure effects on the solute chemical potential. The contribution of solute-solute interactions expected at high concentrations, is uncertain since the partial molar volumes v2o are not known accurately enough. There is evidence, however, that solute concentration effects may play a role in some aqueous systems. Figure 39 shows solubility data on the systems He-HéO and Nz-HZO, both at 25.0°C. On Fig. 39 the quantities plotted are the same as those on Fig 38, i.e. {51(f2/x2P4’) - PVZO/RT w x2. We show curves representing each value of molar volume listed on Table XIV. Although the best fit for He is obtained with v20 = 15.5cm3, this seems too low in relation to other values 209 of the noble gases. 109 P“ 1n (f/xz) — v2° P / RT 12.0 11.8 11.6 11.4 11.2 11.0 210 Y \I L V V l h .‘ ' ' r 1‘ w 4 ,‘ ., . _ u .e He vz" = 15 ems/mole 7 3 a O 4 h ’ He v; = 23 ems/mole - - -( — I . —( ' He v3” - 29 ems/mole ‘ +N3 Vg° - 33 Ema/mole — I— lllllllllllllllllllilllllllll 0 0.001 0.002 0.003 0.004 0.005 0.006 Mole Fraction, X2 Figure 39. anifz/xzi - PVéo/RT z§,mole-fraction solubility for He and N2 at 25°C for various values of vzo. 14. Number Density Scale So far we have analyzed the high pressure results in terms of the mole fraction scale. In this chapter we consider the thermodynamics of solvation on the number density scale. This scale along with its many advantages has been put forth by Ben-Naim.134 To illustrate how this applies to our high pressure solubility data we write the chemical potential (in the T,P,N ensemble) for the solute in the gas and liquid phases: 142' = RT40n(ngA3) + 12:9 (14.1a) (12' = maupz’fi) + 11;" . (14.1b) where A = h/(2xka)“z 1.8 the thermal wavelength of the 14 The first term on the right hand sides solute particle. of Eqs.(l4.la) and (l4.lb) represents the free energy per mole of an ideal gas at the specified solute density, and temperature. The second term, 11:, represents the excess free energy per mole (over the ideal case) due to intermolecular interactions in the gas and liquid phases. Formally u: should also include contributions from the internal partition function of the solute. For simple solutes like the inert gases these contributions are mostly electronic and we assume as usual that they do not change 211 212 when the solute goes into the solvent. The excess chemical potential in phase i, p:i, (i= 1 or g) can be interpreted as avagadros number multiplied by the free energy to take an extra solute from infinity and place it at fixed positions in phase i of the assembled system. (see discussion in Chapter 6 for details.) By equating 1121 with 1129 we find the equilibrium condition in terms of the Ostwald solubility, L, to be It It It Au=p’-pg=-R'ronL, (14.2) 2 2 2 where L = (pal/p29). The key quantity that one wishes to understand in developing a theory of solvation is the excess chemical potential in the liquid, p:l. When the gas phase s is at low density it can be treated as ideal, i.e., p29: 0. l (see For this case Eq.(l4.2) gives a direct measure of p: Eq.(3.11)). When the gas is at a higher density one must account for its nonideality. For the nonideal case we use Eq. (11.2) to find the chemical potential in terms of the solute fugacity. Using the equilibrium condition we now find an expression for p21: *1 92 RT (.12 = -m~m[ —] = «7011(3) . (14.3) f This introduces the new dimensionless solubility parameter, 1 = (pleT / f2). In the low pressure limit 1 reduces to L. Equation (14.3) also holds when the solvent vapor pressure 213 isn’t negligible, as would apply near the critcal temperature of the solvent. In that situation Eq.(ll.1) no longer gives the solute fugacity correctly. Rather one must then employ standard mixing rules in the gas phase to find f .121,l31 2 Because our experimental technique allows for' direct measurement of pa1 we can easily evaluate L and 1. Columns 5 and 6 of Table 13 lists the results of our solubility measurements in terms of the parameter 1. Also included are the values of ”:1 as given by Eq. (14.3). The statistical mechanics of solute-solvent mixtures 135,16 starts with a standard partition function from which, as we have shown in chapter VI, one may obtain the excess chemical potential for a single solute molecule in the liquid solvent14: *1 "2 = -kTOn . (14.4) The key quantitiy in Eq.(14.4), 30’ is the binding energy of a single solute to a fixed configuration of the solute solvent system. The ensemble average of the exponential is taken over all configurations. In practice Eq.(14.4) is difficult to evaluate unless one performs some sort of weighted average which neglects improbable 102,136 1 * configurations. Another way of writing p2 is to use a coupling parameter approach: 214 *1_ 1 pa - Io 0(1)01 , (14.5) where the averaged interaction potential between the solute and the liquid is a function of the coupling parameter, 1.38,39,40 Normally one ignores solute-solute interactions with this technique. If solute-solute interactions become important the resulting chemical potential will be a function of the solute concentration in the liquid, i.e., “:1: F(pzl) 'The solute density can then be determined through Eq.(l4.3). The pressure and concentration dependence of u:l can be evaluated in an manner analogous to our earlier derivation. Starting from Eq. (22) of Ref (115) one can easily show for dilute solutions: -Rron(1) = -RT0n(L°) + Pv2° + pz’ RT( -022 - v20 +RTx° ) +0(p2‘)2 (14.6) where Lois the infinite dilution limit of the Ostwald solubility,and -RT6n(Lb) is the excess chemical potential in the same limit. The number density scale also offers advantages in studying the temperature dependence of the salvation process. The excess molar enthalpy and entropy in the liquid are: €3p*l * 6T 9,. 2 where pzl is given by Eq. (14.3). The excess free energy of solution is commonly described on the mole fraction scale by Henry’s constant via; 1.1;” = RTbn(k") , where k": (fa/x2). In this case the term excess refers to a hypothetical standard state22 and ideal solution.22 Henry’s constant is related to the solubility parameter 1 by: 7=(p,‘+ (02’Hi'l‘lkll (14.88) 4: pilRT/k" , when pzl<