This is to certify that the dissertation entitled The Solubility of Xenon in Simple Organic Solvents and in Aqueous Amino Acid Solutions presented by Jeffrey Frank Himm has been accepted towards fulfillment of the requirements for Ph.D. degree in Physics Mpikggggfl Gerald L. Pollack Date September 3, 1986 MS U is an Affirmative Action/Equal Opportunity Institution 0-12771 RETURNING MATERIALS: MSU Place in book drop to ”BRAKES remove this checkout from n. your record. FINES will be charged if booE is returned after the date stamped below. THE SOLUBILITY OF XENON IN SIMPLE ORGANIC SOLVENTS AND IN AQUEOUS AMINO ACID SOLUTIONS BY Jeffrey Frank Himm A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Physics and Astronomy 1986 c? (“a C‘.\. V97. ABSTRACT THE SOLUBILITY OF XENON IN SIMPLE ORGANIC SOLVENTS AND IN AQUEOUS AMINO ACID SOLUTIONS BY Jeffrey Frank Himm We have measured the Ostwald solubility (L) of 133Xe in a variety of liquids, including normal alkanes, normal alkanols, and aqueous solutions of amino acids, NaCl, and sucrose. For the alkanes and alkanols, measurements were made in the temperature range from 10-50 °C. Values of L were found to decrease with increasing temperature, and also with increasing chain length, for both series of solvents. Thermodynamic properties of solution (enthalpy and entropy of solution) are calculated using both mole fraction and number density scales. Results are interpreted using Uhlig's model of the solvation process. Measurements of L in aqueous amino acid solutions were made at 25°C. Concentrations of amino acids in solution varied from near saturation for each of the amino acids studied to pure water. In all solutions, except those with NaCl, L decreases linearly with increasing solution molarity. Hydration numbers (H), the mean number of water molecules associated with each solute molecule, were determined for each amino acid, for NaCl, and for sucrose. Values of H obtained ranged from near zero (arginine, H-O.2:tO.5) to about 16 (NaCl, H=16.2510.3). to my dad ii ACKNOWLEDGMENTS First and foremost, I would like to thank Professor Gerald L. Pollack, tuna generously provided guidance, encouragement, support, and friendship to make this possible. I would like to thank all of my friends, both in the department and outside of it, whom i have come to know and love over the years. I would like to thank Richard Kennan for his assistance in the preparation of this thesis, and Dan Edmunds for all of his help with troublesome equipment. I would like to thank the Naval Medical Research‘and Development Command (Contract numbers N0001u-80-C-0617 and N0001h-83—K—07N3) and the ExxOn Corporation for their financial support. And finally, I want to thank my family, especially Mom and Dad, for their love and support. 111 TABLE OF CONTENTS LIST OF TABLES.... ........... . ..... ..... ..... .. ........... ..........V LIST OF FIGURES... ........ 0.00.0.0.0...00......OOOOOOOOOOOOOOOOOOOVii INTRODUCTION.OOOOOOOOOOOOOOOOOOIOO0.0.0.0.0.0000...00.0.0000000001 2. THEORYOOOOOOOOOOOOOOOOOOOOOO0.00000000000000000000000.00.00.000005 2 1 IDEAL SOLUTION...........................................6 2 2 RAOULT'S LAW.............................................9 2.3 HENRY'S LAW.............................................11 2.” HARD SPHERES........I...................................13 2 5 26 KIRKWOOD'BUFF THEORY....................................18 APPLICATIONS TO EXPERIMENTS.............................23 2.6.1 ASSUMPTIONS......................................23 2.6.2 THERMODYNAMICS OF SOLUTION.......................2N 2.7 MULTICOMPONENT SYSTEMS......... ...... . ....... ...........27 3. EXPERIMENTAL.OOIOOOOOOOOOOOOOOO0.0.00.0.0...0.0.000000000000000033 SOLUTION COMPONENTS.....................................33 APPARATUS...............................................35 TEMPERATURE CONTROL.....................................37 SHIELDING...............................................37 ELECTRONICS.............................................38 VOLUME DETERMINATION....................................31 SOLVENT PREPARATION.....................................U3 LOADING THE APPARATUS...................................43 SOLUBILITY EQUATION.....................................u5 o omqmszN". wwwwwwwww O 1‘. RESULTS AND CONCLUSIONSOOOOOOOOOO0.00...OOOOOOOOOOOOOOOOOOOOOOO.50 .1 ALKANES O O O I O I O I O O O O O C O O O O O O O I I ........ O O O ...... O O O O O O O I O 50 O 2 ALKANOLS O O O O O O O O O O O O O O O O C O O O O O O O O O ...... O O O ......... O O C . 77 3 AQUEOUS AMINO ACID SOLUTIONS...... ...... ...............107 «Elf-'4: 5. LIST OF REFERENCESOOOO..00...OOOOOOOOOCOOOOOOOOOOOOO0.0.0.0....135 iv LIST OF TABLES TABLE 1. Solubility data for experiments with 133Xe in the normal alkanes. The first row gives the Ostwald solubility measured for each alkane, and the second row is 10 x , where x is the mole fraction solubility of Xe in the normgl alkanes a3 1 atm partial pressure of Xe....................................51 TABLE 2. Chemical potentials in the mole fraction scale, Ana, and in the number density scale, Au°p. The first row is Aug and the second row is Augp. Units age cal/mol......................62 TABLE 3. Entropy and enthalpy of solution in both the mole frap§§on scale and in the number density scale for solutions or xe in the normal alkaneSOOOOOC0.00......0.0.0.00000000000000070 TABLE A. Comparison of experimental binding energies, Eexp’ of Xe in the normal alkanes with binding energies estim ted from the heats of vaporization of the solute and solvent, ESSt.....76 TABLE3§. Ostwald solubility L and mole fraction solubility x2 for Xe in the normal alkanols. The Eirst row for each alkanol is L, and the second row is 10 x2..........................78 TABLE 6.Chemical potentials in the mole fraction scale, Aug, and in the number density scale, Au The first row is Au 09 and the second row is Augp. Units age cal/mol.............?........87 TABLE 7. Entropy and enthalpy of solution in both the mole frac§§on scale and in the number density scale for solutions of Xe in the normal alkanols. (Table 2 of reference 69).........95 TABLE 8. Comparison of experimental binding energies, Eexp’ of Xe in the normal alkanols with binding energies estiBategst from the heats of vaporization of the solute and solvent, ED ....100 TABLE 9. Comparison of calculated and experimental thermodynamics Of SOlUtionOOOOOOOOOOOOOOOOOOO0.0.0.0000...OOIOOOOOOOOOOOC0.00.00.102 TABLE 10. Comparison of calculated and experimental values for the enthalpy and entropy of solution for mixtures of Xe in some of the alkanes and alkanols. Values were calculated using Pierotti's model, and the Lennard-Jones parameters of Wilhelm and Battino.......................................................105 TABLE 11. List of solubilities measured for Xe in aqueous solutions of the amino acids, sucrose, and NaCl. Also included are the solution molarities (M): the solution pH, the volume fraction water, and the hydration number calculated for each experimentOOOOOOOOOOOOOOOOOOOOOOO0.0000000000.0.0.000...0.000.000.0108 TABLE 12. Hydration numbers of the amino acids....................115 vi LIST OF FIGURES Figure 1. Solubility measurement apparatus.........................36 Figure 2. Schematic diagram of the electronics.....................39 Figure 3. Schematic diagram of the degassing system................uu Figure A. Schematic diagram of the gas handling system.............fl6 Figure 5. Normalized counting rate vs. time for a typical experiment(Figure 2 of reference 67). 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...................................................u9 Figure 6. L vs. T for the normal alkanes. The numbers next to the symbols are the number of C atoms in the alkane molecule.......53 Figure 7. x vs. T for the normal alkanes. The numbers next to the symbols are the number of C atoms in the alkane molecule.......55 Figure 8. L vs. n for the n-alkanes at 5 temperatures..............57 Figure 9. x vs. n for the n-alkanes at 5 temperatures.............58 2 Figure 10. L vs. n for the n-alkanes; +, 20°C, our data; A, 25°C, MakranczyE}.floooooooooooooo00000000000000.0000000000000060 Figure 11. x vs. n for the n-alkanes; +, 20°C, our data; A, 2.500,Makganczye_t21.0.0000....0.00......OOOOOOIOOOOOOOOOO0.00.61 Figure 12. Au° vs. T for the n-alkanes. The numbers next to the symbols are tge number of C atoms in the alkane molecule...........6h Figure 13. Au°p vs. T for the nealkanes. The numbers next to the symbols are tge number of C atoms in the alkane molecule...........66 Figure 1“. Au; vs. n for the n-alkanes at'S temperatures...........68 up 2 Figure 16. Entropy of solution for Xe in the n-alkanes.............71 Figure 15. Au vs. n for the n-alkanes at 5 temperatures..... ..... 69 Figure 17. Enthalpy of solution for Xe in the n-alkanes............72 Figure 18. Binding energy, surface energy, and enthalpy of solution for the n-alkanes. (Figure 5 of reference 68).............75 vii Figure 19. L vs. T for the normal alkanols. The numbers next to the symbols are the number of C atoms in the alkanol molecule......80 Figure 20. x vs. T for the normal alkanols. The numbers next to the symbols gre the number of C atoms in the alkanol molecule......82 Figure 21. L vs. n for the n—alkanols at 5 temperatures............8u Figure 22. x vs. n for the n-alkanols at 5 temperatures...........85 2 Figure 23. Au° vs. T for the n-alkanols. The numbers next to the symbols age the number of C atoms in the alkanol molecule......89 Figure 2A. Au°p vs. T for the n-alkanols. The numbers next to the symbols age the number of C atoms in the alkanol molecule......91 Figure 25. Aug vs. n for the n-alkanols at 5 temperatures..........93 Figure 26. Augp vs. n for the n-alkanols at 5 temperatures.........9u Figure 27. EntrOpy of solution for Xe in the n-alkanols............96 Figure 28. Enthalpy of solution for Xe in the n-alkanols...........97 Figure 29. Binding energy, surface energy, and enthalpy of solution for the nralkanols. (Figure u of reference 69)............99 Figure 30. L vs. M for Xe in aqueous amino acid solutions.........118 Figure 31. L vs. v w for Xe in aqueous amino acid solutions.......122 t Figure 32. L/L0 vs. Vt" - HM/55.3u6 for amino acid solutions......126 Figure 33. E vs. molecular weight. ............. ...................129 Figure 3uofiv30 SOIUtion pH......O....00.0.0.0...0.0.0.0000000000131 viii INTRODUCTION The fundamental goal of studies on the solubility of gases in liquids is to understand solution processes from molecular first principles. Given the physical properties of a solvent 1 and a solute 2, we would like to predict the thermodynamics of salvation of 2 in 1. Although this goal is not presently attainable, we have taken a step towards understanding solubility by studying some simple gas/liquid systems. We have measured the Ostwald solubility, L, of 133Xe in a variety of liquid solvents. Ostwald solubility is the equilibrium ratio of the volume concentration of solute gas in the solvent to the concentration 8 2 molecules per unit volume) of solute 2 in the liquid and gas phases, in the gas phase. If 92, p are the number densities (i_e_ the number of respectively, then R/pg. (1) L" D2 2 Xenon was chosen as the solute gas because of its simplicity, because of its importance in the biological and environmental sciences, and because, as will be explained below, it is easy to work with. As an inert gas, Xe interacts weakly with other molecules (short-range, induced-dipole interactions). Due to these weak interactions, the inert gases form prototypical solids and liquids, and have been the subject of many studies."2 3 The general properties of these gases are well known. 133Xe is an inexpensive and easy to detect radioactive isotOpe of Xe. This isotope is of environmental interest since it is a byproduct of 133x6’ as nuclear power reactors. along with Kr, another radioactive inert gas, were the major radioactive contaminants released to the environment during the accident at Three Mile Island.” At a partial pressure of 0.8 atmosphere (atm), Xe is an inhalational anesthetic. The mechanism of anesthesia seems to be the same for all anesthetic agents.5 Although there are indications that the site of anesthetic activity is in the cell membrane, the exact mechanism is not yet known. Cell membranes are, simply put, lipid bilayers in which proteins are imbedded. The building blocks of the bilayer are phospholipid molecules. These molecules contain a polar head group bonded to 2 nonpolar hydrocarbon chains. They are thus amphipathic; polar (or hydrophilic) at one end and nonpolar (or hydrophobic) at the other. Since the cellular environment is aqueous and highly polar, it is thought that hydrOphobic effects may be important in the formation and stability of cell membranes.6’7 The two hydrophobic tails of the lipid molecule are straight chain hydrocarbons, 16-18 carbons long. Thus, the environment inside a cell membrane is similar to that in one of the normal alkanes, i.e. both are nonpolar. This is one of the motivations for using the normal alkanes as solvent. The normal alkanes are a homologous series of straight chain, saturated hydrocarbons. They are nonpolar, implying that interactions with a nonpolar solute will be of the induced-d1pole/induced-dipole type. Physical prOperties of the normal alkanes are well known and can be found in the literature.8 Density as a function of chain length is nearly constant, about 0.7 g/cm’, so that the number of carbon atoms, or CH2 groups, per unit volume is nearly constant at about 3 x 1022/cm3. Thus, a Xe atom in liquid n-dodecane will be surrounded by much the same microscopic environment as one in liquid n-tridecane. The difference is in the number of methyl (CH ) and methylene (CH2) groups with which it 3 interacts and in the geometry of these groups. A second homologous series of solvents studied are the normal alkanols. They differ from the alkanes in the replacement of a hydrogen on a terminal carbon by a hydroxyl (OH) group. Adding the polar hydroxyl group makes the longer chain alkanol molecules amphipathic. It also changes the solvent environment radically. Intermolecular interactions are stronger due to the non-zero dipole moment and due to hydrogen bonding. These stronger interactions show up in increased surface tensions, and in higher melting and boiling points for these liquids.9 We also measured the solubility of Xe in aqueous solutions of amino acids, sodium chloride, and sucrose. In these solvents, we consider complex, three-component systems. Water, in itself, is a highly polar solvent. In addition, amino acids in solution (as the name amino acid implies) become ionized, and have both positively and negatively charged groups. Also, the amino acid side chain can be nonpolar, polar, or charged. By comparing values of L for Xe in amino acid solutions with that for Xe in pure water, we obtain information on both amino acid/water interactions and on Xe/ amino acid interactions. The former can be important in studying protein formation and conformation, while the latter may aid in understanding Xe/protein interactions. Both types of interaction may be important in studies of anesthetic activity. THEORY There are several important relationships that we will need later and which we will now derive. In section 1, we introduce the concept of the ideal solution. All of the gas/liquid systems we studied involved dilute solutions of Xe in the solvent. These systems are known as dilute ideal solutions; i.e. solutions in which the free energy of the solute in solution is the same as that for an ideal solution. In section 2, we derive Raoult's Law, a key prOperty of ideal solutions. We also discuss why Raoult's Law is not applicable to our systems. Henry's Law, which is a generalization of Raoult's Law, is introduced in section 3. We also derive here the relationships between the three quantities KH, x2, and L, which are the Henry's Law constant, mole fraction solubility, and Ostwald solubility, respectively. Section 11 is a discussion of hard-sphere models of solubility. Our results on the solubility of Xe in the alkanes and alkanols will be compared with the predictions of some of these models. Kirkwood-Buff theory is the topic of section 5. It may be possible to use this theory to predict deviations from Henry's Law at high pressures. Section 6 brings together the relationships we will need to analyse our results. Section 7 discusses multicomponent systems, and introduces the concept of hydration numbers. We also derive the equations we will use in our discussion of the Xe/amino acid/water systems. 2.1 IDEAL SOLUTION An important starting point in the study of solubility is the concept of the ideal solution. Ideal solution theory provides a framework with which real systems may be compared. As in the case of a perfect gas, it is the deviations from ideality that yield information on interactions. A < ‘73) moles of A, and Consider two liquids, A and B, at a temperature T. Let V be the molar volume of A (B). If we form a mixture of n A nB moles of B, then A and B form an ideal solution if‘o: (AHm)T’P a 0 , (2) (AVm)T’P - o = v - ”AVA - nBvB . (3) H is the enthalpy, and V is the volume. The subscript m stands for mixing. The first condition, that no heat is evolved or absorbed, indicates that A molecules interact with B molecules as if they were A molecules, and vice versa. The second condition, that no volume changes occur, is necessary to eliminate changes in entropy from that source, since BS 3? Pressure does change with temperature if the volume is held fixed. To prevent changes in entrOpy on mixing, we must assume that the volume does not change, so that as A3 ‘ (BVJT AV = O . The following conditions on the energy of mixing, (mm, and the free energy of mixing, AGm,can also be derived. (AUm)T,P a 0 a (AHm)T,P - P(Avm)T,P (5) and (AGm)T’P a -T(ASm)T,P . (6) Note that (AV)-(-a-AG) -o (7) m 3? m T,n - and AG 2 3 m (AHm) = ‘T [fi-(TJLD n = 0 . (8) We can see that equation (8) is valid by performing the differentiation, (3“) 5T . If we write and using the relation S a - P,n 11A) + ”BUB - 11g) (9) we see (assuming that equations (7) and (8) hold for each component) i :- A,B (10) 'The pi and u; are the chemical potentials of component i in the mixture and in the standard state, respectively. The chemical potential is the free energy per molecule, and the standard state is the pure liquid. Equations (10) have as solution 111 = 112+ RTln(Xi) , (11) so that (ASm)T P a -n Rln(xA) - n , A BRln(xB) , (12) where we have xA and xB for the mole fractions, respectively, of A and B. = 1 - x . (13) 2.2 RAOULT'S LAN Raoult's Law, which we will derive from the assumption of solution ideality, is a simple relation between the mole fraction solubility and the vapor pressure of a solution component. Unfortunately, as we will see below, it is only valid in a few special cases for real systems. If the vapor phases of A and B are ideal, we have (in the gas mixture) 1' j (l/ ) 11 u u 4’ mm P. I:o (1 ) I» where Pi is the partial pressure of i, 111 is the chemical potential of pure 1 in the vapor phase at unit pressure Po. At equilibrium, the chemical potential of‘ i in the gas is equal to that of i in the liquid. Then, * u; + RTln(xi) = “i + RTln(P /P,] . (15) i For pure 1, xi - 1 , P . P° the vapor pressure of pure 1, so that i i ’ N u; - u, + RTln[P;/P,) . (16) Then RTln(x1) - RTln(Pi/P°) - RTln[P;/P,J . (17) 10 01“ P1 - xiP; . (18) Equation (18), which says that the partial pressure of component i in the vapor phase is equal to the mole fraction of component i in the liquid solution times the vapor pressure of pure liquid 1, is known as Raoult's law. There are a few real systems which behave ideally. Most are isotopic mixtures, such as a mixture of 160 and 180 11 2 2’ optical isomers.12 It is not surprising that these systems should be or mixtures of ideal. However, there are other systems, such as the benzene/toluene system at about 90° C, which behave ideally and are not in these two 13 classes. Other possible ideal systems are mixtures of. the rare gases. There are only two rare gas systems, argon/krypton and krypton/xenon, which can be studied over the entire range of composition. Davies_e_t a_l_.1u investigated solutions of argon and krypton at the triple point temperature of Kr (T - 115.77° K). They found deviations of 041% from Raoult's law, and volume changes of less than 1.5% from the ideal over the full range of composition. Calado and Staveley1S found, in a study of Kr/Xe systems, deviations of up to 12% (xKr . 0.3), and changes in volume of about 1.5%. Thus, even when interactions are weak and short range, deviations from Raoult's law are observed. 11 2;} HENRY'S LAW If the temperature is above the critical temperature of one of the components, a more general form of Raoult's law, known as Henry's law, applies. Henry's law states: P = K x. , (19) i.e., the partial pressure of component i is proportional to its mole fraction in the solution. However, the constant of proportionality, KH’ called Henry's constant, is not the vapor pressure of pure 1 as it is in Raoult's law. When expressed in the more general form f a K x (20) where f is the fugacity of 1,7 Henry's law is valid for some gases up i to very high pressures, to 600 atm for N2 and CH” gases in liquid solvents of H 0 and of aqueous sodium chloride.16 2 Henry's coefficient, K is simply related to the Ostwald H9 solubility coefficient L. The mole fraction solubility of 2 in 1 is ) (21) x = n /(n1 + n 2 2 2 where n1 is the number of moles of i in solution. For dilute solutions (x2<< 1) we can write 12 2 1 R - x2. 3“ =[——T—_—+1]‘ (22> LPZV1 + 1 LPZV1 RT where n2 - LPZVI/RT is the number of moles of 2 in volume V1, P2 is the partial pressure of 2, and n1 - 1 is the number of moles of 1 in molar volume V1. Solving equation (22) for P2 we see x RT 2 LV 2 1 For x2 << 1 , P2 = (RT/LV1)x2 (2n) Comparison of equations (19) and (2H) gives 1% = RT/LV1 . (25) Deviations from Henry's law can be related to solvent-solvent, solute- solvent, and solute-solute interactions using the Kirkwood-Buff theory 17 of solutions. Kirkwood-Buff theory is discussed below. 13 2.11 HARD SPHERES Besides the thermodynamic approach to solvation processes offered by ideal solution theory, there is also an approach to solubility from molecular theory. There are two models of liquids which have had some success in predicting the physical properties of liquids and the thermodynamics of liquid mixtures. Both models involve hard Sphere molecules, i.e. molecules with intermolecular potentials of the form (26) where a is the radius of the molecule. The equation of state for a hard sphere liquid is the same for both models, even though they involve different assumptions and methods of derivation. The first model is due to Percus and Yevick,18-22 who developed a collective coordinate description of a system of hard spheres. Percus and Yevick assumed that the pair distribution function was known, and derived from it an integral equation for the system. (The pair distribution function, (r), is the probability that there is aJ 2” molecule a distance r from an 1 molecule.) Thiele23 and Wertheim solved this integral equation, obtaining the equation of state of a hard sphere liquid. Lebowi tzzs later derived the equation of state for a Percus-Yevick mixture of hard spheres. The other model, deve10ped by Reiss, Frisch, Helfand, and Lebowitz,2°'28 is known as scaled particle theory (SPT). In SPT, the pair distribution function for a system of hard spheres is constructed, and from it the equation of state is derived. This equation of state is: 1H 2 _P_.1+v+v ka (27) (1 - y)3 :3p/6 and p is the number density of the hard spheres. As where y = na noted above, this is the same as the equation of state derived from the Percus-Yevick model. 29 3O Uhlig and Eley proposed the following model of the solution process. First, a cavity is created in the solvent which is large enough to hold a solute molecule. Next, the solute particle is introduced into the cavity and allowed to interact with the solvent. The reversible work to create the cavity of radius r is given by W(r) . Anr3P/3 + Anr20(1 - Egg) + Ko (28) where a is the surface tension of the liquid, a is the solvent diameter, 1Q) is a constant which is small compared to the other terms, and 6 is a factor which takes into account the non-zero curvature of the cavity surface. The first term is the work done to create the cavity against pressure P, while the second is a surface energy term. 31 Reiss approximates W(r) for r > a/2, using scaled particle theory, by the polynomial W(r) - k0 + K1r + K2r2+ K3r3 . r > a/2 (29) For r S a/2, the exact expression for W(r) for a hard sphere fluid is W(r) . -kT[ln(1-unr3p/3)] . r S a/2 (30) 15 2 If we require continuity of W, 3% , and-a—g at r - a/2 (so that the Sr function is at least smooth, if not well behaved, at r - a/2) then we 22-2u get the following expressions for the K's 7: I 2 kT[-ln(1-y) + g( 13y ) ] - nPa3/6 x - (AT-)1 6!, + 18( 11', 121+ «Pa? (31) . k: 12y y 2 _ K (2)[1_y +18( 1_y ) ] ZNPa x N J: a "U \ W o By comparing equations (28) and (29) we see that the surface tension can be expressed as: o a kT/(Nna2)[ 13; + 18( yy )2] - 33 . (32) Comparisons of calculated surface tensions with experimental values yield good results (correct order of magnitude) when applied to simple 27 32 liquids, and very good results (i 10%) when applied to fused salts. Pierotti33 derived an expression for the Henry's law constant of a gas in a liquid. He found * 6 1n( ) - -5 33(5—) + -3 + ln(51) (33) KH ' kT kT ° 1 16 The factor c* is an interaction energy which is a function of the polarizabilities and susceptibilities of solvent and solute. The first term is thus related to the free energy of interaction of the solute with the solvent. In the second term, the numerator Cc - W(r12), where W(r) is given by equation (29) with the K's as in equations (31), and + the mean separation is r12 - (r1 diameters are r1 and r2 respectively. V1 is the molar volume of the 1! solvent. The quantity 5 can be estimated from the equation r2)/2. The solvent and solute 6* - (npch/o?2)[a1a2/((u1/x1) + (dz/x2))] . (3“) Here the a are the polarizability and susceptibility, respectively, 1’ xi of component i, m is the mass of the solute 2, and a as given 12" r'12 at above. Using reasonable values for r and r2, and calculating c from 1 equation (311), Pierotti predicted the solubilities, heats of solution, and partial molar volumes of gases in several liquids. The agreement with experiment is good (110% in anH, 120% in 17b, and correct order of magnitude in the enthalpy of solution).33 It is also possible to express 5* in terms of Lennard-Jones parameters. In this form we have3u * 3 ELJ - (2/3)1rpe12012 , (35) where c - (e e )1/2 a - (o + o )/2 and c o are the Lennard- 12 1 2 ’ 12 1 2 ’ i' 1 Jones energy and distance parameters for component i. 17 * , Wilhelm and Battino35 have used ELJ and equation (33) to estimate 6 and a for several gases and solvents. We will use their Lennard- i 1 Jones parameters to predict the thermodynamics of solution for Xe in some of the alkanes and alkanols. Yosim and Owens36 calculated heats of vaporization and fusion for the inert gases, and the heats of vaporization for a number of diatomic and polyatomic gases from the hard sphere equation of state. Agreement with experiment was very good (:51) for most of the monatomic and diatomic molecules, and was good (:10 to 20%) for the larger, polyatomic molecules. Yosim37 calculated heat capacities at the boiling point for these gases. Calculated heat capacities were within 10% of experimental values for the symmetrical gases, and within 301 for the asymmetrical 37 ones. Yosim also predicted the compressibilities of Ar, H N 0 2’ 2’ 2’ and C6H6. Agreement with experiment was good (20%). Why should experimental results compare so favorably with such a simple model? Intermolecular potentials are essentially hard sphere, repulsive, cores with a short range attractive component. Looked at in this way, the attractive component serves primarily to determine the density of the fluid. The density, p, appears explicitly on the left side of equation (27), and implicitly on the right side in the parameter y - npa3/6. So the model does take the attractive component into account, at least in an averaged sense. Longuet-Higgins and Widom38 proposed that a hard sphere liquid be put in a uniform attractive potential, in order to make the attractive component more explicit. Snider and Herrington39 used this idea in the equation of state .3. - 5.2 pm. - x(y) (RT) (36) 18 where x(y) is the right hand side1of equation (27), and A is an 1additional parameter. When generalized to binary mixtures, agreement of calculated values of AHE(excess enthalpy of mixing), ASE(excess entropy' of mixing), and AVE(excess volume of mixing) with experimental values was found to be generally good (:30 to “0%) for 10 equimolar mixtures.39 2.5 KIRKHOOD-BUFF THEORY In studies of solvation processes, and of solute and solvent interactions, we gain information by varying temperature, pressure, and composition of our systems. Temperature variations, as we will see below, yield thermodynamic quantities, such as the chemical potential, enthalpy, and entropy. Variations in pressure and composition, when analysed according to the theory of Kirkwood and Buff, tell us about other properties of the solution. Kirkwood and Buff“0 developed a general statistical mechanical theory of solutions. Using integrals of the radial distribution functions of the different molecular pairs, they derived expressions not only for the partial molar volumes (Vi) and the isothermal compressibility (KT), but also for the osmotic pressure and the derivatives of the chemical potentials (p13). Their equations are applicable to multicomponent systems. For a two component system, it can be shown that17: 19 KT - C/an (37) V1 - [1 + p2(G22 — G12)J/n (38) 62 - [1 + p1(G11 — G12)]/n (39) uij- gig; . uii = r 5% (no) where n ' D1 + D2 + p1°2 0, the gas solubility is less in the salt solution than in water, and salting out occurs. If kS < 0, the gas solubility is greater, and salting in occurs. 56 57 Shoor and Gubbins and Masterton and Lee derived expressions for the Setschenow coefficient from scaled particle theory. Agreement with experimental values of the Setschenow coefficient was good for a variety of gases (He,Ne,Ar,Kr,H 0 N CHA’CZHU’C2H6'SF6) in several 2’ 2' 2’ different salt solutions (NaCl, LiCl, KCl, KI). Masterton58 also used scaled particle theory to predict Setschenow coefficients of 5 gases (He, Ne, Ar, 0 N2) in seawater. The average difference between 2’ observed and calculated values of kS was of the order of 5% at 25°C. Considering the large number of components in seawater, this is very good agreement. Euc ken and Hertzberg.59 studied the salting out of gases in aqueous salt solutions. From their data, Eucken and Hertzberg were able to calculate hydration numbers for each salt in solution. The hydration number of a salt is the number of water molecules associated with each anion-cation pair in aqueous solution. They made several assumptions about the solutions in order to calculate the hydration numbers. The first assumption is that water in an aqueous solution exists in one of two states, either "bound" or "free". Bound water is associated with the salt, and is unavailable to dissolve the solute gas. The rest of the water is free. The solubility of the gas in free water is the same as in pure water. 30 Eucken and Hertzberg also assume that the molar volume of both free and bound water is equal to that of pure water. This model is based on the excluded volume concept. The decrease in gas solubility occurs because of a decrease in the volume of water available, and the gas is excluded from the volume occupied by the salt ions and their hydration shells. Under these assumptions, one can derive the following expression for the hydration number, H, of a salt: L H - (MHZO/M)(vtw - E6) (66) Here, MH 0 is the number of moles of water per liter at temperature T, M 2 is the molarity of the salt solution, vtw is the volume fraction of total water for the solution, L is the Ostwald solubility of the gas in the solution at temperature T, and L is the Ostwald solubility of the 0 gas in pure water at temperature T. We can invert equation (66) and write the solubility in terms of the hydration number H. H (Msalt + MHZO °salt °H20 L - L 1 - M )] (67) 0[ where M1 and p1 are the molecular weight (g/mol) and density (g/l) of i, and M is the molarity. The fraction of the volume which is water, either free water or hydration water, has been written as 31 M v -1-Mi§lE. (68) t" psalt leing equations (65) and (67). we can relate the Setschenow coefficient to the hydration number by: M HM kSM - -log10[1 - M[ salt + ——§19)] (69) °salt °H,o M HM For dilute solutions, i.e., those for which M(__s_a_l_t +-——H-39) (< 1, we salt °H,o can expand the argument of the logarithm to obtain M HM kéM . ( salt + H30)M (70) psalt °H,0 from which M HM k' . ( salt + H30) (71) psalt szo where ké =- 2.303kS . Since ké is of the order of 0.33 l/mol for NaCl solutions, we would not expect equation (70) to be valid for our systems. The Taylor series expansion of ln(1-x) is: ln(1-x) - x + x2/2 + x3/3 + ..., (72) so the fraction error made in neglecting second and higher order terms is of the order of x/2. Since our solution molarities are of the order of 0.5, and ké is of the order of 0.33, we make a 10% error hi 32 neglecting the higher order terms in equation (70). We will therefore use equation (66) to interpret our results. EXPERIMENTAL 3r.1 SOLUTION COMPONENTS 133Xe in trace The solute gas in our experimental system is amounts. The partial [pressure of 133Xe is typically about 1 picoatmosphere (patm) usually mixed with air at 1 atm total pressure. Control experiments were also done in which 133Xe was mixed with naturally occuring, nonradioactive Xe gas at a total pressure of 1 atm. A variety of liquid solvents were used. These included several normal alkanes and alkanols, as well as aqueous solutions of amino acids, NaCl, 133 and sucrose. For the experiments in which the-gas phase was Xe plus nonradioactive Xe, the solvents were degassed before the run, and then 133 saturated with nonradioactive Xe. For experiments in which Xe was mixed with air, the solvents were not degassed. In those cases, the solvents were probably saturated with air which dissolved during production and storage of the solvent. However, this had no measurable effect on our values of solubility. Xenon is one of the noble gases. Its atomic radius is 2.23 A,1 and it is therefore the largest and most polarizable atom of the 5 common inert gases (He, Ne, Ar, Kr, Xe). (A sixth noble gas, radon, has only radioactive isotopes, no stable ones. The longest lived isotope of radon is 222Rn, which has a half-life of 3.8 days.60 It is a product of alpha decay of radium.) The particular isotope of Xe used in our 33 3” experiments, 133Xe, is unstable and decays with a half-life of 5.2145 days.61 It decays by beta emission to an excited state of 133Cs. This nuclear excited state decays60 with a half-life 6.3 x 10.9 sec by eunitting an 81 keV gamma ray. We used the emitted radioactive intensity of these gamma rays as a measure of 133Xe concentration. Xenon-133 was purchased (Medi-Physics, Plainfield, New Jersey) in 20 milliCurie (mCi) aliquots. Typical amounts of 133 Xe used during a run are of the order of 100 pCi. High purity, nonradioactive Xe is commercially available (Matheson, New Jersey, 99.9+%). The normal alkanes (Humphrey Chemical Co., New Haven, Connecticutt) are simple, nonpolar organic molecules. They are "straight" chains of carbon atoms saturated with hydrogen.62 Since they are nonpolar, interactions between alkane molecules, or between an alkane molecule and any other nonpolar molecule, are short range, induced dipole to induced dipole interactions. The carbon to carbon bond length is about 1.5 11,62 so a Xe atom interacts directly with only a portion of a long alkane molecule. 'The normal alkanols (Aldrich Chemical 00., Milwaukee, Wisconsin) differ from the n-alkanes in the addition of a hydroxyl group (Chi) to a terminal carbon.62 Although bond lengths are not affected, the molecule is polar. Thus we would expect stronger interactions between Xe and alkanol molecules than between Xe and alkane molecules. For the experiments involving amino acids (Aldrich), solutions of amino acid in distilled water were made at 25.0 °C. We measured the pH of the amino acid solutions, but made no attempt to control the pH, i.e. no buffers were used since these would inevitably add yet another component to the solutions. 35 Physical properties (density, surface tension, etc.) for the 8.9.63 alkanes and alkanols were taken from the literature, as were the molecular weights of amino acids, and the solubilities of amino acids in water.6u 3 . 2 APPARATUS A diagram of the apparatus is shown in Figure 1. The apparatus can be disassembled at the threaded Joint for cleaning and loading. Two valves, along with several machined brass pieces are soldered together to form the upper portion of. the apparatus, i.e., above the threaded Joint. The bottom portion is a Pyrex container whose mouth is Joined as shown to a glass-to-metal seal. A brass nut is soldered to this seal. The nut and large ball valve form the threaded Joint, which is sealed by wrapping with Teflon tape before assembly. During a run a glass encased magnetic stirring bar runs inside the bottom vessel, which contains the solvent. Commercial stir bars are coated with Teflon. It was necessary to remove this coating since xenon concentrates in it, apparently because xenon is highly soluble in Teflon. We encase the stir bar in glass so the bare stir bar will not corrode in the experimental solvent, which is acidic in some cases. Teflon is used in two other places; in the threaded Joint as tape, as noted above, and in the solid packing material in the ball valve. We iglore any contribution of these seals to the measured solubility. In both cases, the surface area of Teflon exposed to Xe is small. This, in light of the very slow diffusion rate of gases in solids Detector L Pb Shield“). \. _ luelte Window mu. w'v. —_# . . ~ . Bel Volvo / ‘ '\' . H Pb Shield“). "read“ Joint - q t \. 1 ' 'Vhb Gloss to Metal Seal ‘ Brass Nut ( V Vee = th-vhb we” Figure 1. Solubility measurement apparatus. r Stir Bar 37 9 cmz/sec), leads to a negligible contribution. Also, we have (D - 10' never seen any evidence of a Xe sink, which would have shown up in either leak tests or dilution runs. 3;}TEMPERATURE CONTROL Temperature was controlled to 10.1°C (Lauda/Brinkmann K-2/RD circulator) both during a run, and in making up experimental solutions. The bath fluid was an approximately 11:1 mixture of water:ethylene glycol. The ethylene glycol decreases the problem of freeze-up of the float in the circulator. 3 . 11 SHIELDING We used Pb shielding of thickness 5 1 cm to reduce background radiation and to prevent emissions due to 133Xe decays outside of volume V“) from reaching the detector. One cm of Pb attenuates 81 keV gamma rays by a factor of 106.6:5 Shielding effectiveness was checked by putting ”’Xe in V with the apparatus and shielding in place. In rest these cases, we did not observe anything above the background rate. 38 3 .5 ELECTRONICS A diagram of the electronics is shown in Figure 2. Gamma rays of energy 81 keV are detected with an NaI(Tl) scintillation detector (Harshaw, type 7SF8). A high voltage power supply (Power Designs, Inc. model 15700 supplies 1100 volts for the photomultiplier via the pre-amp (Ortec model 276). Detector output passes through an amplifier (Ortec model 1185), then through a single channel analyser (SCA, Ortec model 1406A). The window on the SCA is set to exclude energies below about 10 keV and above about 90 keV. The window was set by feeding signals of known amplitude into the amplifier, and observing the output on a multichannel analyser. The resulting output was then compared to a 133Xe spectrum. Two timers (Ortec model 719) are externally supplied a 0.1 sec timing pulse by a minicomputer (Micro DevelOpment T001 1000). The computer is programmed to take data and print out the results. It counts the number of decays in A00 sec, makes suitable corrections to the total, then prints out the result. It then waits an additional 200 sec before counting again. Each hour the average of the previous six values is printed out. Two corrections are made to the data. Due to the nature of the electronics, a pulse into the detector is not counted at the computer if it too closely follows an earlier pulse. This "dead time" can be measured by putting signals of a known frequency into the electronics, and observing the output. In this way, we determined the dead time for our system to be about 1 ~ 3.33 x 10"6 sec. 39 HIGH VOLTAGE POHER SUPPLY GIGNGL TRIGGER COMMEBTIONG Figure 2. Schematic diagram of the electronics. MO Radioactive decay is a random process, and consequently,tflm probability of k events, given a mean of u events, is given by the Poisson distribution66: k P(k) - —‘-‘-— . (73) kieu P(k) is the probability of k events per unit time, give a mean of u events per unit time. If we let T, the dead time, be our unit of time, and u be the mean number of events per time T, then the probability that a single signal occurs within time 1 of an earlier signal is given by 9(1) . ° . ne'“ . (7n) 1!eu The mean number of signals per unit time 1 is p, and is typically of the order of 2 x 10-3, so we may use the approximation: P(1) ~ u(1 -‘u) ~ u . (75) The actual counting rate, Nact' can therefore be approximated by Nact a Nobs(1 + u) ’ (76) where Ni, is the observed counting rate. We have programmed the be computer to make this correction to the data. The second correction which must be made to the data is to remove excess counts due to natural background radiation. We measure background ”1 radiation several times over periods of several hours with the apparatus and shielding in place, but without any 133Xe in the apparatus. Background is subtracted after the dead-time correction has been made. Typical backgrounds for a run are about 1100 counts in NOD sec. Counting rates due to 133Xe vary from about 250,000 counts/1100 sec at the start of a run, to about 50,000 counts/H00 sec at the end of a run. 3.6 VOLUME DETERMINATION Volumes are indicated in Figure 1. There are seven volumes which Vm 8 and Vee’ are the same for every run. The remaining two, V must be determined for each run. Five of them, , V V V hb’ are rest’ sb’ (2) and V (1) determined at the beginning of each run. The gas volume V8 2! is the volume above the ball valve which initially contains all of the 133Xe. (2) The gas volume V8 is the remaining gas volume. The volume of the liquid solvent is V The volume of the glass encased stirring bar is £0 vsb and vhb is that of the small hole in the ball of the valve. The volume below the ball valve is designated by vee’ and vrest is the total (1) volume excluding volume Vg . To determine vee (ee stands for everything else), we weigh the apparatus before and after filling Vee with water at a known temperature. We make a buoyancy correction for the weight of air displaced by the water. If mo is the observed mass, then the true or bs actual mass, m is: act ’ ”2 P m air,T] p H20,T /[1 ' e (77) act ' mobs where p1 T is the density of i at temperature T. Vee is then to v ”—53“— . (78) 9° pH20,T Vhb (hole in ball) was calculated from measurements made on a disassembled ball valve. For V , rest vrest ' vee + vhb ' (79) The volume of the glass-encased stirring bar, V8b , was determined by measuring the water displaced when the stirring bar was immersed in a known volume of water in a graduated cylinder. (1) We determined the initial gas volume V8 , by using a dilution (1) measurement. Xenon-133, initially confined to volume Vg at a concentration of C1 , is allowed to expand throughout the volume V21) + Vrest . If the final concentration, corrected for decay of 133Xe is Cf , we define the ratio a to be: Cf vé1) a . _ ' (17" g (80) 1 vg + vrest from which v“) - —-9‘—- v . (81) g 1 - a rest ‘13 3.7 SOLVENT PREPARATION Alkanes (99% pure) and alkanols (purities: methanol,99.9%; propanol and butanol, 99+%; pentanol, octanol, decanol, and undecanol, 99%; hexanol, heptanol, and dodecanol, 98%; nonanol and tetradecanol, 97%; 200 proof ethanol) were used as purchased. Amino acid solutions were prepared in 250.0 1 0.1 cm3 volumetric flasks in the following way. To a predetermined mass of amino acid, we added distilled water to fill the flask to within 1 cm3 of the 250.0 cm3 mark. For experiments involving Xe atmospheres, the water was first degassed by boiling under vacuum (Figure 3), then was saturated with Xe under a Xe pressure of 1 atm. The amino acid and water were also mixed under a Xe atmosphere. After temperature equilibration at 25.0°C, enough water was added to bring the total solution volume to 250.0 cm3. The outside of the flask was‘ carefully dried, then the flask and solution were weighed, with the usual buoyancy correction being made for the weight of displaced air or IXe. Solution density was calculated, along with the volume fractions of both water and amino acid. A small sample, about 5 cm3 , of the solution is used to determine the pH (Brinkmann pH meter with combination electrode). 3.8 LOADING THE APPARATUS (1) With the ball valve closed, air is evacuated from volume Vg through the Hoke valve. A trace amount of 133 into V(1) 8 Xe is allowed to expand , and then air or Xe is let in to bring the pressure to 1 atm. NM to Vacuum Pump Figure 3. Schematic diagram of the degassing system. ”5 (1) When using nonradioactive Xe, V8 was loaded using the gas handling system shown in Figure A. The Hoke valve is then closed. Approximately 220 cm3 of solvent is poured into the glass portion of the apparatus, and is weighed. The gas handling system of Figure 14 was also used to load the solvent under a Xe atmosphere. The buoyancy corrected weight divided by the solvent density yields the volume of the (2) liquid solvent, V2. The secondary gas volume, V8 , is: (2) _ _ V8 - vrest V8b V1 . (82) After putting in the stirring bar, the apparatus is assembled. Six to seven turns of Teflon tape are used to seal the threaded Joint. Scratch marks on the ball valve and on the brass nut ensure that it is assembled reproducibly. 3.9 SOLUBILITY EQUATION Solubility is calculated using conservation of 133Xe. Initially, (1) 133Xe is confined to volume Vg at a concentration of C1 .When equilibrium has been reached, the 133Xe is distributed throughout (1) (2) volumes V , 8 8 amount of 133Xe in the solvent is Cvai’ solubility of 133Xe in the solvent, and C (1) 8 and V as well as throughout the solvent. The total where L is the Ostwald f is the equilibrium concentration of 133Xe in volume V .Then 146 3 13 To Vacwm PUMP 3.“ Valve Pot Loading Xe 1' ...... r—1 Gauge Solvent Beauvoir Nonraoioeotlve No my Promo Fitting Comeote to upper Portion ot Apparatus ' 0 Valve Lower Portion of WING Figure A. Schematic diagram of the gas handling system. "7 C V(1) . C eAAt(V(1) + V22) 1 g f 8 + Lvll . (83) 133 where )1 is the decay constant for Xe ( A - 0.9177 x 10-” min-1), and At is the time elapsed between measurement of C and C . The term em“; 1 f 133Xe atoms. We can also see that LV is a correction for the decay of A is the "effective gas volume" of the liquid solvent. Rearrangement of equation (81) gives (1) (1) (2) c v v + v L = file Mt—é— - 3 v 3 . (811) r 9. 9. Equation (8A) is valid for measurements at a single temperature. Experiments involving the alkanes and alkanols were done at a variety of temperatures. The temperature dependent form of equation (8A) is: Ci -AAtv(1) v(1) + V(2) L(T) = -C—-e -%—a(i~) - 3 V 3 3(1) + 1 - R(T) (85) f 9. 2. where R(T) - p(T)/p(T2) is the ratio of the solvent density at temperature T to the solvent density at temperature T the loading 2" temperature. For the alkanes and alkanols, T was usually taken to be 9. 20.0°C, except for those which were solid (heptadecane, octadecane, nonadecane, and eicosane) at that temperature. In these cases, T was 2 taken to be either 30°C or 110°C, depending on the melting point of the alkane. During a run, C1 is measured hourly for six to eight hours. Each value is corrected for decay, and the mean value is calculated. After the gas is uniformly distributed in Vé”, as determined from the N8 observed decay rate, the main valve is opened and the stirrer is Started. We calculate the Ostwald solubility, L, each hour. WhenI. reaches a constant value, equilibrium has been reached. Time to equilibrium is typically 12 hours. We wait an additional 8 to 12 hours to be sure of equilibrium (see Figure 5. Figure 5 appears as Figure 2 in reference 67). The equilibrium values of L are averaged to determine the Ostwald solubility. For amino acid experiments, a second pH sample is taken at the end of the run. A9 50" r I I I I I I I I I I —-40 u: I r l I 1 l I l I I- {.2 A. c°° . -1 :1 , - >30 . - b 0 E " . ' '2: . 1% - ._ 7. °. . I 6 . ° .— .s .. 4 .. lmXele"At larb Time [hours] Figure 5. Normalized counting rate vs. time for a typical experiment(Figure 2 of reference 67). 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. RESULTS AND CONCLUSIONS The results for Xe solubility in the normal alkanes and normal 68,69 alkanols have been previously published. In addition, some of the results on the solubility of Xe in aqueous amino acid solutions have also recently appeared.70 ii . 1 ALKANES We measured the Ostwald solubility, L, of 133Xe in normal alkanes at five temperatures in the range from 10°C to 50°C.68 Values of L, and the mole fraction solubility, x2, are listed in Table 1. Mole fraction solubilities were calculated using equation (21). Figures 6 and 7 are plots of L versus Temperature (T) and x versus T, respectively, for the 2 normal alkanes. Figures 8 and 9 show L versus n, the number of carbon atoms, and x versus n, respectively, at five different temperatures. 2 The Ostwald solubility decreases with both increasing T and increasing n, while the mole fraction solubility decreases with T, and increases with n. Thus, even though the number of Xe atoms per unit volume decreases with increasing n, the number of Xe atoms per alkane molecule increases with n. The decrease of Xe solubility with increasing temperature is typical of gas/liquid systems. 50 51 TABLE 1. Solubility data for experiments with 133Xe in the normal alkanes. The first row gives the Ostwald solubility measured for each alkane, and 2, where x2 the normal alkanes at 1 atm partial pressure of Xe. the second row is 102x is the mole fraction solubility of Xe in Temperature 10°C 20°C 30°C “0°C 50°C Number of carbons in n-alkanes 5 L 6.u1:o.12 5.A8:0.11 x 3.03 2.56 6 5.91:0.03 5.07:0.10 u.55:o.o3 3.18 2.68 2.36 7 5.u110.03 u.67i0.07 u.1310.07 3.7510.07 3.37:0.0 3.26 2.77 2.u1 2.15 1.90 8 n.9910.03 h.36i0.05 3.9010.02 3.A7i0.05 3.3110.06 3.3” 2.86 2.51 2.20 1.95 9 A.70:0.02 n.1uio.o3 3.70:0.03 3.32:0.0M 2.99:0.03 3.95 2.99 2.62 2.30 2.0“ 10 u.u2 3.9210.03 3.52:0.0A 3.1h 2.8V 3.59 3.08 2.71 2.38 2.11 11 n.1810.ou 3.7210.03 3.35:0.ou 3.00 2.71 3.62 3.16 2.79 2.A5 2.18 12 9.03 3.5910.02 3.22:0.01 2.9010.01 2.6a 3.76 3.28 2.88 2.55 2.28 13 3.88:0.0 3.uuto.o1 3.0910.02 2.80 2.53 3.87 3.37 2.96 2.63 2.3a 1h 3.76:0.01 3.3510.02 3.02:0.01 2.72 2.N9 u.oo 3.fl9 3.08 2.73 2.uu 52 TABLE 1 (cont'd.). Temperature 10°C 20°C 30°C N0°C 50°C Number of carbons in n-alkane 15 L 3.2“:0.02 2.9210.01 2.69 2.91 x2 3.59 3.17 2.81 2.51 16 3.19:0.01 2.85:0.03 2.5710.0 2.35:0.0 3.67 3.26 2.89 2.59 17 2.7610.03 2.51i0.01 2.30 3.3” 2.98 2.68 18 2.7110.01 2.u7i0.01 2.2510.01 3.A5 3.08 2.76 19 2.9210.01 2.21:0.01 3.17 2.8” 20 2.3610.0 2.17:0.0 . 3.29 2.92 4.5 Ostwald Solubility ‘0” 4.25 4.00 3.75 2.75 Ostwald Solubility' + I I I I + 5 .. A .A a El 7 m '1’ o e " V 9 o A - ‘V El A d o ‘7 El _‘ 0 gr C! o - v 0 v — l L, l l ' L to so so 40 so eo ' Temperature (°C) , .l. I r I I -+ :o - A A :4 B D :2 q o 4' <> :3 V A V 14 .7 C3 +_ _ o 9' AA 1 a + 8 A _ m + 3' a ‘ 1) .— 1 l l l l :0 so so 40 so so Temperature (°Cl Figure 6. L vs. T for the normal alkanes. The numbers next to the 53 symbols are the number of C atoms in the alkane molecule. Ostwald Solubility 511 r l I l I 4- 15 D 17 A to :e 3.00 - ‘7 19 ,_ 4. Al 2.75 - E; ‘- 4. 2 50 £5 . 1- % + 1 - g _ 2.00 l l l l l 0 10 20 30 40 50 60 Temperature (°C) Figure 6. (cont'd.). 55 3'50 v I I I I 6 + 5 A 3 U 7 + . O G 0 1: 2.73 G _ .3 P A v 4.1 8 2.50 r- + o _ t a v 2.25 - - 3 B o 1: 24m+- g - ‘.75 L l l l I 0 :0 ~20 so 40 so 30 Temperature (°C) 4°25 I I T r F 4.00 — v + to .. o 0 A 11 o 3.75 - El El :2 - ('1 A 0 13 x sAm- + v ‘V “ - c: o D 3.25 h- E T °'" A m s. 8 L A ll. 2.75 " + V "" m 8 H 2.50 - v - g t 2.25 F- a d 2.00 l l l E f 0 io 20 so 40 50 so Temperature (°C) Figure 7. x vs. T for the normal alkanes. The numbers next to the symbols are {he number of C atoms in the alkane molecule. Mole Fraction x 100 3.75 I 56 A r I I + 3.50 '- o «- [3 3.25 '- A o —I -P ‘7 <> 3.00 " a d + 15 A ' £5 16 .p ‘7 2J3I' c] 17 To - 0 1! B V 19 A 2.50 "' + d 2.25 I 1 1 1 ' I 0 10 -20 30 40 50 Figure 7. (cont'd.). Temperature (°C) 57 8'5 4: I I I I I I I I I I I47 I I I I °°° "' + + 10°C ‘ >, _ A 20°.c . o r-o 5.0 _ A 4- ° ‘°.,° - 44 4- ‘7 50 C a 4 5 - 3 A " 3 ° A + ._ 0 E] A A + :3 3.0 VI <9 ‘9 c] ‘7 <9 - C] c] '5 25 v v V 3 ° ° 0 a . "' v 0 o '1 2.0 "- ‘ 1.5 I L I, I, l I ,L, I I I L, I I I, I J 4 5 B 7 a 9 10 11 12 13 14 15 1B 17 1a 19 20 21 Number of Carbons. Figure 8. L vs. n for the n-alkanes at S temperatures. 58 Mole Fraction x 100 1-5 I I I I I I T I I I I I I I II 4.0- + - + + + AA + A A a + E! o + A El 3.0-+ AA an 0 v- 0 A alt! .0 v ‘ BB 00. V + 10°C 0° 7V A 20°; 0 777 u 30%: 2.0- v7 o 40°04 V 50°C 1.5 I LII I LglJl I III II 455789101112131415161718192021 Number of Carbons Figure 9. x2 vs. n for the n-alkanes at 5 temperatures. 59 71 Makranczy _e_t_ ii. measured the Ostwald solubility of Xe in the normal alkanes n'pentane through n-hexadecane. Their data are plotted, along with our data for 20°C, in Figure 10. The mole fraction solubilities are shown in Figure 11. It is interesting to note the differing behavior of x as a function of n. The data of Makranczy_e_t 2 El- show a decrease in the mole fraction solubility with n. 72 Clever gt a_1_. measured the solubility of He, Ne, Ar, and Kr in the normal alkanes n-hexane through n-decane, n-dodecane, and n-tetradecane at several different temperatures. Their data show a decrease in x2 with with increasing n for He and Ne, but Ar shows an initial decrease in x2, followed by an increase for the longer chains. The mole fraction solubility of Kr in the n-alkanes increases with increasing n. This would lead you to believe that the mole fraction solubility of Xe in the normal alkanes should increase with increasing n. In table 2 are values for Aug. calculated using equation (58), 09 2! function of temperature, and as a function of n, are shown in Figures and Au calculated using equation (61). The chemical potentials as a 12, 13, 111, and 15. A least squares fit of the A112 versus T data for both Au; and Augp, yields straight lines (r 2 0.999) for each of the alkanes. It is from the slope and intercept of these straight lines that we obtain the entrepies and enthalpies of solution. The entrOpy of solution, obtained via equation (59), is the negative of the slope of the best fitting line, and the enthalpy is the intercept on the ordinate axis. Values of the entropy and enthalpy are listed in Table 3, and are plotted versus n in Figures 16 and 17. 60 I I I I I I I I I + This Work 5'5 " + A Mokronczy 0t. 01. ‘ > :I': 5 o '- t q H or! 5 4- n 4 — - 3 A -+ H + C3 .. .- (D 4.0 A + .9 E 3.5 - A + + - ID 4' 4. 35 so A "' m ' A . C) At 2.5 P- 45 .4 A 20 I 4 I I I I I I I I 9 H, 4 5 6 7 8 9 10 11 12 13 14 15 1B 1 Number of Carbons Figure 10. L vs. n for the n-alkanes; +, 20°C, our data; A, 25°C, Makranczy 33 El‘ Mole Fraction x 100 61 4'00 T I I I I I I I I I I I + This work 3'75 - A Nakranczy at. 21. + A + 3.50 — + q 4. 335r- + - .p + 3.00 +- + -I + 2J5- A -t _ i! £8 24m _ +' 9* A ,3 A, A A 2.25 - .1 2.00 I I I L I I I I I I L I 4567891011121314151817 Number of Carbons Figure 11. x vs. n for the n-alkanes; +, 20°C, our data; A, 25°C, Makganczy _e_§ _a_1_. 62 TABLE 2. Chemical potentials in the mole fraction scale, Ap°, and in the number density scale, Au°p. The first row is An; and the segond row is Augp. Units are cal/mol. Temperature 10°C 20°C 30°C 90°C 50°C Number of carbons in n-alkanes 5 Au; - 1967:8.9 213618.? Aug” - -10u5:8.u -991:8.7 6 1991 2109 225719.0 ' “999 “995 “91319.0 7 1926 2089 2296 2390:9.3 2596t9.6 “950 “898 “859 -82219.3 “78019.6 8 1913 2071 2218 2376 2529 “909 “858 “820 “779' “769 9 1899 2095 2195 ' 2397 2998 “871 “828 “788 “797 “703 10 1879 2027 2173 2327 2977 “836 “796 “758 “712 “670 11 1867 2012 2157 2307 2958 “805 “765 “728 “689 -690 12 1897 1991 2136 2283 2929 ' -73u ~7uu -709 -662 -623 13 1830 1976 2120 2269 2911 “763 “720 “679 “691 “596 19 1811 1955 2096 2291 2383 “795 “709 “666 “623 “586 63 TABLE 2 (cont'd). Temperature 10°C 20°C 30°C 90°C 50°C Number of carbons in n-alkane 15 Au; .p 1939 2080 2223 2366 Aug - “685 “695 “609 “565 16 1926 2062 2209 2395 -667 “631 “587 “599 17 2098 2186 2325 “612 “573 “535 18 2027 2166 2306 “601 “563 “521 19 2199 2286 “550 “509 20 I 2139 2270 “539 -998 69 259° I I I I I 2500 I- “F 5 “ .A 0 . A 2400 P a 13 " H O 0 17 E3 2300 r- .— "\ 9 F. g 2500 r- - o m 2100 - " “ 3. q 2000 - ,1 “ 1900 *- .— I‘ I I L I .00 1 0 1° 20' 30 4° 50 50 Temperature (°C) 2000 I I I I 4. 2500 '- A D .. 2400 - o H C: E§"2300 “- ‘\. '3 2a” 3 b ON 2100 1- 3. ‘<3 2000 " 1'00 " 1500 ‘ l J 0 10 20 so 40 50 50 Temperature (°C) Figure 12. 00° vs. T for the n-alkanes. The numbers next to the symbols are tie number of C atoms in the alkane molecule 65 2800 I I I I I amor- -+ 0 ' A 12 -, 2400 - D ‘° . '4 :3 0 20 r 2 a... ' \ o '3 £3 zm0- ‘ ‘ o o 01 2100 r- ' é a 2000 r- ' 1000 - n ' 1000 ~ ‘ l l L J 0 10 20 30 40 so 00 Temperature (°C) I I I An: (cal/mol) l 0 10 20‘30 ‘0 5° 3° Temperature (°C) Figure 12. (cont'd.). 66 In" I I I I I Z: O E \ H I! 3 8a 3. d -1000 - + + - l l l, l I - o . Temperature (°C) A1129 ° (cal/mol) l l l 1, l '1100 0 10 20 :0 40 50 Temperature (°C) Figure 13. Au°p vs. T for the n-alkanes. The numbers next to the symbols are t a number of C atoms in the alkane molecule 67 A1129 (cal/Illa 1) -1100 Apt?” “(cal/mull -1000 '1100 Figure 13. I _l l l ' l 20' 30 40 so 00 Temperature (°C), I I I I I J I I 10 (cont'd.). 20 30 40' so Temperature (°C) 68 2400 2100 Aug (cal/mall 1900 lllJJll 4 5 B 7 B 9101112131415181‘718192021 1800 Number of Carbons Figure 1”. Au; vs. n for the n-alkanes at 5 temperatures 69 "00 I I‘ I I I I I I l' I I I* I I I I -500 '- V v V d V 0 f... -800 "' v V 0 a a - ‘5. V 0 ° D B A A H -700 r- V 0 a a A A .1 u 3 A + v V D A + '300 F- + -I ‘0» C1 45 -F <> 40°C ' -1000 - A + V mac 4 4. _1‘00 I .L I II I II Iiil I I I I I L I I 4 5 B 7 B 9 10 11 12 13 14 15 1B 17 1B 19 20 21 Number of Carbons 09 Figure 15. A02 vs. n for the n-alkanes at S temperatures 70 TABLE 3. Entropy and enthalpy o{3§olution in both the mole fraction scale and in the number density scale for solutions or Xe in the normal alkanes. Hole Fraction Scale Number Density Scale Enthalpy of Entropy of Enthalpy or Entropy of Temperature Number or Solution Solution Solution Solution Range of Carbons AH; (cal/mol) ASS (cal/moi K) Aflgp (cal/mol) As;-p (cal/moi K) Experiments 5 “2818 t 200 “16.90 t 0.69 “257“ t 200 “5.90 t 0.69 10 “ 20°C 6 “2691 t 132 “16.37 t 0.U5 “2230 t 132 “H.35 t 0.US 10 “ 30°C 7 “2H15 t 61 “15.36 t 0.20 “2120 t 61 ~u.1u t 0.20 10 “ 50°C 8 “2929 t 77 “15.3“ t 0.26 “2120 t 77 “4.29 t 0.26 10 “ 50°C 9 “2378 t “5 “15.09 t 0.15 “2050 t “5 “H.16 t 0.15 10 “ 50°C 10 “2351 t 51 “13.93 t 0.17 “2010 t 51 -u.16 t 0.17 10 “ 50°C 11 “2308 t 55 -1u.7u t 0.18 “1970 t 55 “H.11 1 0.18 10 “ 50°C 12 “2278 t 2H “1u.56 t 0.08 “1930 t 2H “H.0“ t 0.08 10 “ 50°C 13 “2278 t 32 “1H.51 t 0.11 “1930 t 32 “9.13 t 0.11 10 “ 50°C 1" “2239 t 23 “1“.30 t 0.08 “1860 t 23 “3.93 t 0.08 10 “ 50°C 15 “2233 t 30 “19.23 t 0.10 “1860 t 30 —u.01 t 0.10 20 “ 50°C 16 “2180 2 b0 “19.00 t 0.13 “1830 1 “0 “3.97 t 0.13 20 “ 50°C 17 “215“ 1 102 “13.68 t 0.33 “1770 t 102 “3.8“ t 0.33 30 “ 50°C 18 “2191 t 38 “13.91 t 0.12 “1810 t 38 “3.99 t 0.12 30 “ 50°C 19 “2152 t 97 “13.73 t 0.31 “1820 t 97 “0.07 t 0.31 N0 “ 50°C 20 “2117 3 23 “13.58 t 0.08 “1690 t 2“ “3.68 t 0.08 “0 “ 50°C 71 0 I I I I I T I I I I I I I I I I -2 — - .. -« - A A A A AAA A A A A A A A A A.- X -5 .. £5 .— H o \ '3 -10 - A A529 --4 u v -12 ’- .1 > '1‘ "' + + + + + a- + + + '1' '1' '1 ° + + + + t; + m '16 _ —d -20 L I I 14 I I I I I L J I I L I 4 5 B 7 B 9 10 11 12 13 14 15 18 17 1B 19 20 21 Number of Carbons Figure 16. Entropy of solution for Xe in the n-alkanes Enthalpy (cal/moi) “1500 “1700 “1800 “1900 “2000 “2100 “2200 “2300 “2400 -2500 “2300 “2700 -2800 “2900 -3000 72 .+ AH? A AH"?D + I l l l. L IA 1. l I II L *— 4 l l l l 5 B 7 B 910111213141515171819 Number of Carbons Figure 17. Enthalpy of solution for Xe in the n-alkanes 73 From Figure 16, we see that both As; and 1339 are negative, and increase slowly with n for n S 8. We can interpret this in several different ways. One explanation is that the Xe has more freedom of movement, or is less constrained in solution, for larger n. A second interpretation is that the Xe has a greater disordering effect on the liquid molecules for large n. The actual situation is probably a combination of these two. Clever73 has calculated the entropy of solution of Xe in n-hexane, n-dodecane, and isooctane. He obtained values of “15.92, -15.511, and -15.51 cal/(mol K), respectively. The agreement with our results is good. Clever gt 11,72 calculated the entropy of solution for He, Ne, Ar, and Kr in the normal alkanes n-hexane through n-decane, n-dodecane, and n-tetradecane. All exhibit nearly constant As; (He, “10.“ 1 0.7; Ne, “10.8 1 0.“; Ar, “13.0 i 0.0; and Kr, -13.7 i 0.2 all in dimensions of cal/mol K). For Xe in the same seven solvents, we determined ASE - “15.13 :I: 0.25 cal/mol K. This is about the value we would expect if we view the inert gases as a homologous series. g and AHEp, are negative and generally increase with increasing n. This is the proper qualitative Like the entropies, the enthalpies, AH effect we would expect according to the cavity formation model of 29 30 [Htlig and Eleyu Assuming the Xe/alkane interaction is nearly constant as a function of n, we would expect this increase in AH, since the surface tension of the alkanes increases with chain lenght, and it is more difficult to create a cavity in a liquid of higher surface tension. 73 calculated AHg for Xe in n-hexane, n-dodecane, and isooctane. His values are -2582, “2273, and ~2089 cal/moi, respectively. Clever 711 Our values for these same quantities are -2691 :1: 132 for n-hexane and “2278 :1: 211 for n-dodecane, taken from Table 3. Our value for n-octane, “21129 :I: 77, also compares favorably with his value for isooctane. The agreement with our data is good. Figure 18 is a plot of the three quantities in equation (6H) for the alkanes n-pentane through n-hexadecane. The binding energy shown has been calculated from the enthalpy, AHE, uwrgY1NA. As expected, the binding energies are nearly constant, varying by a few percent from the mean of ”500 cal/mol. The gradual increase in and the surface energy, Eb with n is consistent with the stronger interactions between Xe and alkane at large n. A plot similar to Figure 18 can also be made for A1151). However, the difference in enthalpies, AH°p - AH° is nearly constant, about 350 2 2' cal/mol, and the behavior of E will be much the same. b Table u lists values of E calculated from equation (6“), and the b binding energy estimated as the geometric mean of the heats of vaporization of Xe and the alkanes at their normal boiling points. Agreement is good, within 20%, for all of the alkanes shown. The estimated binding energy also shows an increase with increasing n, much as the values obtained from experiment. However, the estimated values increase at a greater rate. 75 5°°°IIIII I I I I I I I El E1 4500 +— 0 E, :1 13 El '3 a q E! D [g :2 4000 - ' + - Enth5lpy -1 g A Surf5c5 En5rpy \ 3500 .. D Binding En5rgy _ H (D 3 3000 - - > + + 2500 - A q co 4- + A ‘- + + A 4* 4‘ 7 + 00 A A + C 2000 - A -4 LIJ A £5 1500 - A _ 1000 I I I I I I I I I I I I 4 5 6 7 B 9 10 11 12 13 14 15 18 17 Number of Carbons Figure 18. Binding energy, surface energy, and enthalpy of solution for the n-alkanes. (Figure 5 of reference 68). 76 TABLE u. Comparison of experimental binding energies, Eexp, of Xe in the normal alkanes with binding energies gggimated from the heats of vaporization of the solute and solvent, Eb . Number of EexP Best b b Carbons per n-alkane (cal/moi) (cal/moi) 5 u261 1071 6 0306 H307 7 0226 M510 8 0373 A701 9 MN33 R860 10 uu9u 5026 11 u526 5167 12 “558 5299 13 “615 5u1u 1A 0628 5539 15 M668 5636 16 N651 1 5751 77 ”.2 ALKANOLS We also measured the Ostwald solubility of 133Xe in the normal alkanols at five temperatures in the range from 10°C to 50°C. The values we obtained for L and x are listed in Table 5. 2 Figures 19 and 20 are plots of L versus T and x2 versus T. As with the alkanes, both L and x2 decrease with increasing temperature. 7H Komarenko and Manzhelii measured the solubility of Xe in n“propanol in the temperature range from “80°C to -30°C. They found that the mole fraction solubility decreased with increasing temperature, not only for Xe in n-prOpanol, but also for a variety of gas/alcohol systems. Solute gases were spherically symmetric (Ar, Kr, Xe, CH , and u CF”) and solvents included methanol, ethanol, n-prOpanol, and n-butanol. The behavior of L and x2 as a function of. n is more complicated for the alkanols than for the alkanes. Figure 21 shows L versus n at five different temperatures. The sharp initial increase, followed by the slow decrease, is probably due to the change from the highly polar environment of methanol to one which is becoming more alkane-like for large n. Figure 22 shows x versus n at five temperatures for Xe solubility in the normal alkanols. As with the alkanes, the number of dissolved Xe atoms per alkanol molecule increases with increasing n. The mole fraction solubility also has its greatest rate of change at small n, when the liquid environment is presumably changing most rapidly. 78 TABLE 5. Ostwald solubility L and mole fraction solubility x2 for ‘3’Xe in the normal alkanols. The first row for each alkanol is L, and the second row is 102x2. Temperature 10°C 20°C 30°C “0°C 50°C Number of carbons in n-alkanol 1 L 2.“6 2.20:0.01 1.98:0.02 1.79 x 0.“23 0.370 0.325 0.288 2 2.79:0.00 2.“?20.02 2.22:0.00 2.02:0.02 1.8510.01 3 3.0210.01 2.6510.05 2.38:0.06 2.16 1.9810.01 “ 3.0“ 2.6810.01 2.“0 2.17 1.98 1.17 1.01 0.88“ 0.782 0.699 5 2.97 2.6210.00 2.36 2.1310.00 1.95 1.35 1.17 1.02 0.903 0.809 6 2.97 2.61:0.02 2.3“ .12 1.92 1.55 1.33 1.17 1.0“ 0.920 7 2.91 2.5710.01 2.31 .09 1.91 1.73 1.“9 1.31 1 16 1.03 8 2.86 2.52:0.03 2.2510.02 2.05 1.88 1.89 1.62 1.“2 1.26 1.13 9 2.79 2.“9t0.00 2.2“ 2.0“ 1.77 1.55 TABLE 5 (cont'd.). 79 Temperature 10°C 20°C 30°C “0°C 50°C Number of carbons in n-alkanol 10 L 2.7“ 2.“3¢0.02 2.20 2.00 1.83 x2 2.19 1.89 1.67 1.“9 1.33 11 2.3“10.01 2.1110.01 1.92 1.76 1.98 1.7“ 1.55 1.39 12 2.12:0.02 1.9“:0.01 1.78 1.89 1.68 1.51 1“ 1.91:0.01 1.76:0.01 1.91 1.72 80 - A + 1 > 5.00 a A s 4-1 U 7 :2 edsh- ° <0 5 "1 A V 11 g B '3 2.50 - + o (D V 3 73°. 2 00 v a ' o P A 1.: "' v a C1 + 1.75 r- 17 1.50 l l J l I 0 1o .20 so 40 . s0 Temperature 1°01 3°25 r I I f - I + 2 , 2.00 p 3 A 4 . j: I: 5 «u 4- <> 12 ,., 2.7s - dd 45 a £3 £3 2.50 " 4. 8 8 2.25 '- 4. u w. o 8 ‘3' 2 00 r- '1' 1; ' ° 9 o 1.7s - ° .so ' ' . 1 ‘ ' 1 0 10 20 so 40 so Temperature (°C) Figure 19. L vs. T for the normal alkanols. The numbers next to the symbols are the number of C atoms in the alkanol molecule. 81 + 5 3.00 r- T ,. + 1A 5 t: A D 10 ..., 2.7s - 13 ° ‘1 ‘ 4-0 a X H 2.so - a “ O + a) u 2.2s 1- e '1 .3 +- 3 2.00 L 6 "‘ +a - o + m o o 1.7s - ° " 1.1s0 L ' L ' L 0 10 20 so 40 so 00 Temperature (°C) Figure 19. (cont'd.). 82 + 1 V .A 2 a 1.28 - . D 5 ‘ g V o 4 V 3 x 14m - ° 7 “ a. C: 49 17 .3 can a o» v ‘6 ' A '3 ° 0! ‘ ‘5 E) L A '3 u. 0.50 - A . 7 CI 4- 4. A“ r4 4L .* g 0.. - - 0.00 I I. I l I 0 1o 20 so 40 so 50 Temperature (°C1 . + 5 CI 0 "' '3 o 5 S + a o 4-0 . (J m E! I. + A “- 1.8 - El .. + A .2 + 9 E 1.00 - A “ 4. ‘,:73 l l I I ' I 0 1o 20 so 40 so 50 Temperature (°C) Figure 20. x vs. T for the normal alkanols. The numbers next to the symbols gre the number of C atoms in the alkanol molecule. -M01e Fraction x 100 1.75 1.50 1.25 1.00 83 Temperature (°C) Figure 20. (cont'd.). l T I F I + T A d + 3 ° A -I + B o -+ 10 A to 2 t: c. .. . W 0 14 + I I I I l 10 20 30 40 so so 8h 3°50 I I I I I I I I I I I I I I + 10°C 3.2: h- A 20°C -+ 5. + D 30°C . 3.00 - + o 40°C _ .3 + + + V so°c ..., + .g 2.75 - + A 4' + .. 3 2.30 - + A a _ A A A .. '0 ° ° 0 o m 3 o o o <-> B CD 4> ‘V’ ‘7 1.75 - V V v .. 1.50 I, I _I I I I I I .I j, I I I I o 123400700104112431415 Number of Carbons Figure 21. L vs. n for the n-alkanols at 5 temperatures. Mole Fraction x 100 85 235 I I I I I I r I I I U l I a00+~ +' A - + .A c1 0 1.75 - A .. + A E! B 0 7 14m - ‘* A ‘3 o G’Iv . . -P A; [3* , ‘7 V? ijflr- '3 o - + A El 9 V 1.00 r + A I3 ‘9 V +' 10°C A A B e V A 20:0 __ <9 ‘7 E] 30 C °°75 4. El v o 40°C V A $ v 50% 04m- 8 _ 0.25 J I I I I‘ I L I I L I L L I 0 I 2 a 4 a a 7 a 9 i0 11 12131415 Figure 22. x 2 Number of Carbons vs. n for the n-alkanols at 5 temperatures. 86 7h Komarenko and Manzhelii observed similar behavior for solutions of Kr in methanol, ethanol, n-propanol, and n-butanol at a variety of temperatures. Table 6 also contains Aug and A1459 for the Xe/alkanol systems. °p respectively, as a function of T. 2 9 Figures 25 and 26 show the same chemical potentials as a function of n. Figures 23 and 2” show Au; and Au As with the alkanes, least squares fits to the Au versus T data yield straight lines with slope -AS and intercept AH. Entrepies and enthalpies of solution are listed in Table 7 (Table 2 of reference 69), and are plotted versus n in Figures 27 and 28. We see the same type of transitional behavior in the Au versus n graphs as we saw in the solubility versus n graphs (Figures 21 and 22). It would be interesting to extend these measurements to higher temperatures and larger n to see if the Xe/alkanol systems become more like the Xe/alkane systems. 133 Entropy of solution for Xe in the alkanols is very close to that of Xe in the alkanes. The behavior of both As; and ASE,p as functions of n are the same as for the alkanes, and are about 10% larger in magnitude. We conclude that the Xe is either more constrained, or has less of a solvent disordering effect (or more of a solvent ordering effect), in the alkanols than in the alkanes. 7” a value of 75 We obtain from the data of Komarenko and Manzhelii, AS° - 49.114 cal/mol K for Xe in n-propanol at -55°C. Abraham lists 2 the entropy of solution for Ar, Kr, and Rn in methanol (-16.0, -17.lI, and -22.14 cal/mol K) and Ar and Rn in ethanol (-15.1 and -18.1 cal/mol K) at 25°C. These compare favorably with our values for Xe in methanol and ethanol (-18.82 and -17.87 cal/mol K, respectively). 87 TABLE 6. Chemical potentials in the mole fraction scale, Au°, and in the number density scale, Au°p. The first row is Au° and the se and row is 09 2 2 Au2 . Units are cal/mol. Temperature 10°C 20°C 30°C 10°C 50°C Number of carbons in n-alkanol 1 3076 1 8 3263 i 9 3152 1 9 3610 t 9 Au - 50;" - '507.1¢8.1 -160.138.7 -111.5:9.o -362.319.3 2 2800 2983 3163 3311 3515110 ~577.9 -527.9 -181.5 ~136.0 -395.1:9.6 3 2619 2800 2975 3116 3316 -621.3 ~567.7 -521.3 -178.9 ~137.1 1 2501 2677 2819 3019 3187 -625.8 -571.7 ‘ -527.1 -182.16 -139.0 5 2121 2591 2760 2929 3093 -613.3 -561.3 -516.5 -169.7 -127.2 6 2313 2515 2680 2811 3011 -611.9 -559.1 -512.7 -167.0 -119.6 7 2281 2152 2611 2776 2936 -601.6 -591.1 ~503.3 -157.5 -113.5 8 2231 2101 2563 2721 2877 -590.7 -537.3 -189.6 -115.8 -1o1.1 9 2191 2351 2509 -577.5 “530.5 -181.5 TABLE 6 (cont'd.). 88 Temperature 10°C 20°C 30°C 10°C 50°C Number of carbons in n-alkanol 10 Au° p - 2151 2311 2161 2618 2773 in; - -567.8 -518.1 -175.5 -132.0 -388.1 11 2286 2111 2593 2716 -196.2 -119.8 -106.3 ’363.8 12 2392 2512 2691 -153.5 -110.8 -369.6 11 2165 2609 -101.0 -361.8 89 Aug (cal/mol) 2200 I I I 3L1 I 0 10 20 30 40 50 Temperature (°C) 320° I I I I I 3000 - 2000 - l 2400 - Aug '(caI/mol) 200° 3 I J L I I 0 10 20 30 40 50 Temperature (°C) Figure 23. Au° vs. T for the n-alkanols. The numbers next to the symbols age the number of C atoms in the alkanol molecule. 2400 Aug (cal/mol) Figure 23. r I I I I + 10 p A 11 1:1 :2 o u /,, ///////i I I I I 10 20 30 40 50 Temperature (°C) (cont'd.). A029 (cal/moi) Apgp .(callmolI Figure 21. Au the symbols a é ééééé 2 6 ’2 2 I 91 I I I I, 10 20 30 40 50 00 Temperature (98) I I I I I I I I I to 20 so 40 so 00 Temperature (°C) ‘9 vs. T for the n-alkanols. The numbers next to go the number of C atoms in the alkanol molecule. 92 ‘350 I F I I ..400 .- :3 3 «so P "\ I". 8 -um b 8‘.» :1, -550 - < 4mo- 450 J I I I 0 10 20 30 40 Figure 21. (cont'd.). Temperature (°C) 93 3700 5 I I I I I 1 1 l I I l l l 3500 - D V 'I' 10°C -1 A 20°.c .. ° 7 [:1 30°C q :3 3300 A o 40% O E] 0 V V 50°C H A B 0 V d! 2900 L. <> ‘7 ‘7 ‘_ 3 + A 13 o m 0 v v i + A .B a 0 e V G 2500 h- + A 13 ° .. 2300 r- + + A A A .. + «1- 2100 I I I I I I I ‘I' I I I I ll 0123453709101112131415 Number of Carbons Figure 25. Au; vs. n for the n-alkanols at 5 temperatures. 91 ’3“ I I I I r I I I l fi I I I I -sso - - <> ‘7 my ‘7 '000 '- v v -1 ::: C1 ‘7 ‘7 " <> 4) <> 0 <-> v v V 0 E5 1-450 '- <> C] E] -1 \ A o o e E! '3 soo 1:1 0 o ' E' A u - - + 13 E] q A. A E] 13 B A A A -550 - A - :L + 4. +. + 10% _,_ + + , a. + Cl 30 c '550 '- 0 40°C .— ‘7 50°C _700 L 4 L L L I L I I L I L I L 0123456789101112131415 Number of Carbons °p vs. n for the n-alkanols at 5 temperatures. Figure 26. A02 95 TABLE 7. Entropy and enthalpy 0(33olution in both the mole fraction scale and in the number Xe in the normal alkanols. (Table 2 of reference 69) density scale for solutions of Mole Fraction Scale Number Density Scale Enthalpy of Entropy of Enthalpy or Entropy of Temperature Number of Solution Solution Solution Solution Range of Carbons Aflg (cal/mol) As; (cal/mol K) AHEp (cal/mol) ASEp (cal/mol K) Experiments 1 -2250 1 120 -18.82 1 o.uo -1880 1 120 -u.8u 1 0.h0 10 “0°C 2 -2257 1 88 -17.87 1 0.29 -1869 1 88 -u.57 1 0.29 10 50°C 3 ~2300 1 88 -17.39 1 0.29 -1910 1 88 ~u.57 1 0.29 10 50°C 3 -23u5 1 88 -17.13 1 0.29 -19u3 1 88 -u.66 1 0.29 10 50°C 5 -2331 1 88 -16.79 1 0.29 ~192u 1 88 -u.6u 1 0.29 10 50°C 6 -236H 1 88 -16.6u 1 0.29 ~19S9 1 88 -u.77 1 0.29 10 50°C 7 -2317 1 88 -16.26 1 0.29 -1922 1 88 -u.67 1 0.29 10 50°C 8 ~23ou 1 88 -16.05 1 0.29 -1900 1 88 -u.6u 1 0.29 10 50°C 9 -2310 1 180 -15.90 1 0.62 -1890 1 180 ~M.65 1 0.62 10 30°C 10 -223u 1 88 ~15.50 1 0.29 -1826 1 88 -u.u5 1 0.29 10 50°C 11 -2210 1 130 -1S.3S 1 0.u1 -1790 1 130 -u.u1 1 o.u1 20 50°C 12 -21h0 1 210 -1u.96 1 0.66 ~1720 1 210 -u.2o 1 0.66 30 SO’C In -2ouo 1 “30 -1u.uo 1 1.30 ~1630 1 R30 -3.90 1 1.30 no 50°C 96 ° IIIIIIIIIIIIII —2I— u-I A - _ - Ad 3“ AAAAAAAAAAAA 9.. c, .3 .. .— 5 .4 H "P + AS? 3-“... - v A AsgP >-sa- - '3 c: -18 P- 4. '+- 'F 4- ‘- UJ .5 4' 4' 4. -1e- 4- - + . _20 I Ll LI L I Ll LL! l l 0123455739101112131415 Number of Carbons Figure 27. Entropy of solution for Xe in the n-alkanols. 97 "500 I I I I r I I I I I I I I A o A o E ‘\. AL Al £5 r4 -1900 r- 45 AL -4 (D A A A 3 A -2000 - _ 3., -F 2' -ano - - IO '4- I: té-amop- 4. ~ In '* + + '-amo + + - T" + + + +- 4400 L L L I L T I I L L I I I I O 1 2 3 4 ‘5 B 7 B 9 10 11 .12 13 14 15 Number of Carbons Figure 28. Enthalpy of solution for Xe in the n-alkanols. 98 The transitional behavior of the chemical potential as a function of n is thus due to the enthalpic contribution of the solvation process. This contribution is shown in figure 28. The initial decrease, followed by the increase, of enthalpy versus n contrasts with the Xe/alkane systems, where AH is an increasing function of n. Our value of AH; - ~2250 cal/mol for Xe in methanol is 75 intermediate in value to the data listed by Abraham of -1170 and -3820 cal/mol for Kr and Rn in methanol, respectively. The data of Komarenko and Manzhelii7u yield AH; - ~2758 cal/mol for the system Xe/n-propanol at -55°C. Our value is -2300 cal/mol for the same system at 25°C. The chemical potential in the mole fraction scale is dominated by the entropic contribution for the alkanols, Just as for the alkanes. Here, TASE is about twice AHé. The number density scale yields the opposite conclusion, namely that the solution process is enthalpically dominated. Figure 29 is a plot of the three quantities in equation (6“) for the alkanols methanol through n-dodecanol. The binding energy Eb was calculated from AHS and unrgY1NA through rearrangement of equation (6“). The average binding energy of the Xe/alkanol systems is about 31 higher than that of the Xe/alkane systems, and is due to the greater surface energy contribution. Table 8 lists both our experimental E and Eb b’ estimated from heat of vaporization data, for the normal alkanols. As with the alkanes, estimated values agree to within 201(fi’the experimental values. Here, too, the estimated values increase with n at a greater rate than the experimental values. 75 Abraham has compiled values of the standard free energy, enthalpy, and entropy of solution for many gaseous nonpolar 99 550° I I I I I I I I I I I l 5000 i- B E '4 E3 .. a £3 E] B F1 4500 *- C] .a g '3' '3 + em 1 ‘ . DY E 4000 P A Surface Energy ‘0 El Binding Energy 13 mmo- 1 > E, 3000 '- -‘ a) ‘5’ £5 ‘5 lu + .1 4- t :1 " 4- + A '+ + 2000 - A A .. 1500 L I I I I I I I I I I L 0 1 2 3 4 5 5 7 8 9 10 11 12 13 Number of Carbons Figure 29. Binding energy, surface energy, and enthalpy of solution for the n-alkanols. (Figure A of reference 69). 100 TABLE 8. Comparison of experimental binding energies, Eexp’ of Xe in the normal alkanols with binding energieseggtimated from the heats of vaporization of the solute and solvent, E b . exp est Number Eb Eb of Carbons (cal/mol) (cal/mol) 1 AZ?” ”708 2 “271 “990 3 nu32 5182 h R628 5266 5 R650 5358 6 ”721 5579 7 5562 8 "777 5H89 9 14853 5911! 10. “831 5658 101 nonelectrolytes in a variety of nonaqueous solvents and in water. He found that the data could be correlated using an equation of the form: AP; . £8 + d (86) where P stands for the free energy G, enthalpy H, or entropy S. R is a parameter, close to the solute's radius in A, characteristic of the solute gas, and i and d characterize the solvent. Table 9 lists the values of AGE, Aha, and A83 at 25°C, calculated from Abraham's parameters for the solvents shown, using a value of 2.16 for R for Xe. The agreement with our experimental values is good. Calculated free energies of solution are about 10% too high for all solvents. The calculated enthalpies of solution are all within 101 of experiment, except for 1-decanol, which is 180% too small in magnitude. The agreement between calculated and experimental entropies is excellent, within 5% for both the alkanes and alkanols. We can calculate the thermodynamics of solution from Pierotti's 33 model. From equation (33) above: Gc Gi RT ln(KH) - ET + ifi" + ln(Vf) , (87) where 102 TABLE 9. Comparison of calculated and experimental thermodynamics of solution. calculated Experimental 0 0 0 O 0 0 A02 AH2 A82 AG AH2 A82 cal cal cal cal cal cal SW9“ (EST) (5:31") (mmol x) (as: (as: (“—1.01 x) hexadecane 2200 -2210 -1fl.7 199“ -2180 -1H.00 decane 2210 -21uo ~1H.5 2100 -2351 -1h.93 hexane 2230 -2330 “13.7 2183 -2691 -16.37 1-decanol 2710 -1580 -15.5 2398 -223fl -15.50 1-octanol 2720 ~2090 -16.8 2M82 -230M -16.05 1-butanol 2930 '2590 '17.9 2763 -2395 '17.13 1-propanol 3070 -2170 -17.7 2883 -2300 -17.39 ethanol 3220 ‘2180 “18.1 3073 '2257 -17.87 methanol 3500 '2120 ‘18.8 3358 '2250 '18.82 c 12 2 _ 12 ‘fif ' 6(1_y)[2(;:‘) (FT-)] 2 1‘122 r'12 1 + 1817¥§) [(FT—J ‘ (r1 ) + 3] ‘ ln(1-y) (88) and 0 e i _ 12 3 fi- 3.5551Ip(--—kT )r12 . (89) If we compare equations (25) and (87), we see that -1nL-—‘=‘+—. (90) We have for the free energy of solution in the mole fraction scale RT ' 2 - -RTln[(E-v-5-§— *1)1] . (91) Au; - -RT1nx 1 2 The entropy of solution is given by equation (59). o.j_ 0 AS 8T(A"2) . (59) 2 P,n r and e are independent of It is easy to show, assuming that r 12, 12 1! temperature, that 1ou G - 1 - - l._$ As; - Rlnx2 + RT(x2 1)1T up (up + TJRT “P Go ' T=§Lfif + ln(1-y) ' Y i" l" .1. 2 .13 2 - .12 1 +181,.,) [(,1 ) (,1 1+ ,1) . <92) where up is the thermal expansivity of the solvent, and Aug . -RT1nx2 + TAS§ . (93) 35 and Using the Lennard-Jones parameters of Wilhelm and Battino, equations (92) and (93) , we can calculate the entropy and enthalpy of solution for some of the alkanes and alkanols. These calculated values are listed in Table 10, along with our experimental values for the same solvents. The agreement with experimental values is very good (1101), except for AH; for tetradecane (1251). Agreement is not as good for the two alkanols. Calculated entropies and enthalpies differ from experimental results by 20% and 50%, respectively. The deviation of the theoretical values from the experimental results occur in the regions we might expect. For the alkanes, the largest deviations occur for the longer molecules. The free energy of cavity formation, 0c, is calculated assuming the solvent molecules are hard spheres. Since the normal alkanes are decidedly non-spherical, such deviations are to be expected. Methanol and ethanol also show large differences between calculated and experimental values. We believe the source of these 105 TABLE 10. Comparison of calculated and experimental values for the enthalpy and entropy of solution for mixtures of Xe in some of the alkanes and alkanols. Values were calculated using Pierotti's model, and the Lennard-Jones parameters of Wilhelm and Battino. Calculated Values _ Experimental [glues Enthalpy of Entropy of Enthalpy of Entropy of Solvent Solution Solution Solution Solution AH; (cal/mol) ASE (cal/mol K) AH; (cal/mol) A85 (cal/mol K) Hexane -2517 -16.57 -2691 “16.37 Heptane —2u95 -16.38 -2A15 -15.36 Octane -2932 “16.22 ~2fl29 “15.3“ Nonane -2337 -15.97 -2378 -15.09 Decane -196u -15.60 -2351 -1u.93 Dodecane -2056 -15.61 ~2278 -1u.56 Tetradecane -1738 ~1H.90 -2239 ~1h.30 Methanol -3195 -22.76 -2250 -18.82 Ethanol -3158 -21.60 -2257 ~17.87 106 differences is also the free energy of cavity formation term. The difficulty in this instance, however, is probably the strong dipole/dipole interactions of the alkanols. Because of these interactions, they are not "hard" spheres. There are several possible directions in which this work might be extended. First, further studies should be done on Xe solubility in the normal alkanes and alkanols at higher temperatures, and with longer molecules. The goal here would be to see if the thermodynamics of the Xe/alkanol systems becomes more alkane-like for large n. It might also be interesting to study the solubility of Xe in solvents consisting of a normal alkane "doped" with one of the alkanols. A second course of research would be to extend these studies to the solubility of Xe in other homologous series of nonpolar solvents, specifically the halogenated normal alkanes such as the perfluoroalkanes and.perchloroalkanes. These liquids have stronger intermolecular interactions than the alkanes, as evidenced by increased melting and boiling points. However, these higher boiling points would permit solubility measurements to be made with the shorter, more spherical molecules. A third direction for further research is to modify the hard sphere theory. Modifications would include the addition of an attractive potential to the hard sphere equation of state, similar to thatcn’ Longuet-Higgins and Widom.38 This would take account of strong interactions between solvent molecules. A second possible modification of the theory might take into account the non-spherical shape of a solvent molecule. 107 L3 AQUEOUS AMINO ACID SOLUTIONS ‘We measured the Ostwald solubility of 133Xe in aqueous solutions of NaCl, sucrose, and 20 amino acids at 25°C. Some of these results have 70 Values of L obtained are listed in Table 11. recently been published. Also included in this table are the solution molarity (M), the solution pH, the total volume fraction of the solution that is water (vtw)’ and the hydration number H as calculated from equation (66). Our values for the mean hydration number for each amino acid are listed in the column at the left of Table 12. Our method for calculating v w has been described in section 1.7 1: above. The values of Vt" can also be calculated from the solution molarity M, and the partial molar volume vpm of the amino acids, via the equation: vtw - 1 - Mvpm/1000 . (87) Using equation (87), and partial molar volumes of the amino acids at 25°C given by Lilley,76 we can calculate the volume fraction water. Agreement with values obtained using our method is good. For several of the amino acids, we were not able to obtain good values for the hydration numbers. These include isoleucine, leucine, phenylalanine, tyrosine, and tryptophan. The ILE data seem to fall into two different groups, with mean values of 13.3 and 5.8. We were unable to obtain a consistent value for this amino acid, since its solubility in water is too low. 108 TABLE 11. List of solubilities measured for Xe in aqueous solutions of the amino acids, sucrose, and NaCl. Also included are the solution molarities (M), the solution pH, the volume fraction water, and the hydration number calculated for each experiment. Volume Ostwald Solution Solution Fraction Hydration Amino Solubility Molarity pH Water Number Acid L M v H tw Alanine 0.09h 0.N50 6.2 0.973 10.60 (Ala) 0.090 0.661 6.55 0.973 9.21 0.088 0.873 6.35 0.9fl6 7.3“ m.w. 89.09 0.088 0.900 6.1 0.9“” 7.00 0.076 1.350 6.27 0.917 8.20 0.079 1.500 6.00 0.908 6.00 0.070 1.800 7.03 0.889 7.03 Arginine 0.100 0.N38 10.6 0.9u5 0.20 (Arg) 0.100 o.u39 11.3 0.995 0.20 0.093 0.919 10.8 0.88“ 0.fl0 m.w. 17N.20 0.09“ 0.920 11.h 0.88“ -0.17 0.103 0.300 10.6 0.963 -1.60 0.095 0.600 11.1 0.925 2.65 0.098 0.760 11.0 0.905 -1.u2 o.o9u 0.760 11.0 0.905 1.33 Glycine 0.082 1.0u3 6.35 0.953 9.52 (Gly) 0.092 0.500 6.50 0.978 12.18 0.085 1.066 6.0 0.953 7.85 m.w. 75.07 0.089 1.071 6.96 0.952 8.2“ 0.075 1.500 6.2 0.93“ 8.36 0.065 2.130 6.96 0.903 7.53 0.068 2.131 5.8 0.9ou 6.82 0.06u 2.132 6.uu 0.903 7.77 109 TABLE 11 (cont'd.). Volume Ostwald Solution Solution Fraction Hydration Amino Solubility Molarity pH Water Number Acid L M v H tw Hydroxyproline 0.09“ 0.5u8 6.0 0.953 6.69 (Hyp) 0.097 0.5u9 5.9 0.953 3.82 0.083 1.156 5.96 0.901 5.65 m.w. 131.13 0.086 1.160 5.8 0.901 n.28 0.078 1.589 5.9 0.86M h.u6 0.070 2.173 5.7 0.813 3.89 0.069 2.3 5.6 0.790 3.35 0.068 2.399 6.0 0.797 3.66 Lysine 0.09“ 0.510 9.9 0.9”5 6.32 (Lys) 0.088 0.511 9.5 0.9“6 12.5h 0.079 1.07 9.7 0.885 7.23 m.w. 1H6.19 0.082 1.09A 10.0 0.880 5.38 0.053 2.19 9.9 0.763 6.65 0.06u 2.19 10.3 0.756 3.8V 0.056 2.775 10.5 0.687 3.17 0.05h 2.800 10.3 0.687 3.51 Sucrose 0.097 0.250 0.9h8 7.28 0.090 0.M98 0.893 n.88 m.w. 3h2.3 0.085 0.750 0.8“1 2.89 0.075 0.998 0.788 u.u6 0.062 1.500 0.679 3.97 0.052 2.000 0.570 2.20 0.0u6 2.250 0.516 2.02 110 TABLE 11 (cont'd.). Volume Ostwald Solution Solution Fraction Hydration Amino Solubility Molarity pH Hater Number Acid L M vW H Proline 0.101 0.300 6.25 0.976 ”.27 (Pro) 0.100 0.600 6.29 0.950 0.61 0.098 0.600 6.26 0.950 2.35 m.w. 115.13 0.092 0.779 6.50 0.935 n.77 0.096 0.900 6.19 0.925 1.19 0.093 0.900 6.12 0.926 2.99 0.091 1.199 6.18 0.901 1.96 0.090 1.200 6.23 0.900 1.76 0.086 1.600 6.2“ 0.867 1.93 0.088 1.600 6.18 0.867 1.27 0.078 2.098 6.25 0.825 2.35 0.082 2.098 6.11 0.825 1.36 0.073 2.88fl 6.26 0.758 1.33 0.071 2.88“ 6.08 0.758 1.69 0.069 3.151 6.15 0.737 1.51 0.065 3.u22 6.88 0.711 1.58 0.063 3.518 -- 0.703 1.71 0.065 3.518 6.2” 0.705 1.uu Serine 0.09“ 0.383 5.9 0.977 13.0u (Ser) 0.101 0.171 6.0 0.990 12.03 0.099 0.277 6.13 0.983 9.80 m.w. 105.09 0.095 0.383 6.1 0.977 11.67 0.101 0.171 6.3 0.990 12.03 o.1ou 0.085 6.1 0.995 9.03 111 TABLE 11 (cont'd.). Volume Ostwald Solution Solution Fraction Hydration Amino Solubility Molarity pa Hater Number Acid L M Vt H w Threonine 0.070 1.uuu 6.0 0.887 8 69 (Thr) 0.090 0.680 6.1 0.9u7 7 97 0.090 0.680 6.1 o.9u7 7.97 m.w. 119.12 0.101 0.300 --- 0.977 u.u6 0.08u 1.060 5.9 0.918 6.56 0.095 o.u9o 5.9 0.962 7.u3 0.081 1.250 6.15 0.903 6.15 0.083 0.870 5.2 0.933 9.5a 0.098 0.300 5.6 0.977 9.68 0.072 1.uuu 5.3 0.887 7.96 0.105 0.150 6.5 0.988 ---- Cysteine 0.097 0.509 5.75 0.962 5.10 (Cys) 0.080 1.158 n.93 0.913 7.57 0.097 0.509 6.0 0.963 5.21 m.w. 121.16 0.081 1.158 5.8 0.91u ' 7.16 0.091 0.680 5.2 0.9“9 7.37 0.098 0.350 u.8 0.97“ 7.82 0.086 1.000 5.5 0.925 6.29 0.088 0.8u0 5.7 0.938 7.10 0.102 0.180 5.2 0.9870 7.61 TABLE 11 (cont'd.). 112 Volume Ostwald Solution Fraction Hydration Amino Solubility Molarity Water Number Acid L M v H tw Valine 0.098 0.3h3 0.968 7.01 (Val) 0.097 0.3uu 0.969 8.67 0.103 0.153 0.986 5.17 m.w. 117.15 0.102 0.158 0.985 7.96 0.099 0.3uu 0.969 5.6V 0.102 0.155 0.986 8.h8 0.100 0.250 0.978 7.66 0.1036 0.080 0.9931 10.9 0.1036 0.080 0.9932 10.93 Histidine 0.100 0.259 7.91 0.97“ 6.5a (His) 0.103 0.126 7.80 0.987 6.72 0.100 0.259 7.25 0.97” 6.5u m.w. 155.16 0.102 0.126 7.25 0.988 11.30 0.101 0.192 7.32 0.981 8.12 0.1025 0.126 7.88 0.9878 9.1a 0.103 0.060 7.15 0.99” 20.57 0.1031 0.0601 7.75 0.99u7 17.68 Isoleucine 0.100 0.192 6.h1 0.979 10.26 (Ile) 0.102 0.09u3 6.78 0.990 16.28 0.099 0.192 6.2 0.980 13.27 m.w. 131.18 0.105 0.0h70 6.3 0.995” 5.69 o.1ou 0.0939 6.2 0.9905 5.5 0.1027 0.1u3o 6.6 0.98u6 6.09 0.101u 0.1921 6.73 0.9802 6.80 0.1031 0.1u3o 6.69 0.9853 n.92 113 TABLE 11 (cont'd.). Volume Ostwald Solution Solution Fraction Hydration Amino Solubility Molarity Water Number Acid L M vt H w Glutamine 0.098 0.219 5.0 0.979 13.77 (Gln) 0.102 0.100 5.30 0.990 16.35 0.100 0.160 “.8 0.985 1“.39 m.w. 1“6.15 0.0989 0.219 5.0 0.9801 11.90 0.10“1 0.0501 6.68 0.9960 15.36 0.1033 0.0999 5.7 0.9910 9.1“ 0.095 0.219 “.6 0.980 21.17 0.100 0.100 “.7 0.991 26.35 0.102 0.050 5.“ 0.996 31.12 Methionine 0.101 0.215 5.8 0.977 6.22 (Met) 0.103 0.103 5.9 0.989 9.30 0.101 0.160 5.6 0.98“ 10.78 m.w. 1“9.21 0.1000 0.215 6.17 0.9772 8.70 0.1038 0.1029 6.35 0.9895 5.5“ 0.100 0.103 5.7 0.989 2“.50 0.103 0.050 5.9 0.995 25.79 0.098 0.215 5.7 0.977 13.51 0.1036 0.0501 6.78 0.9955 20.00 Asparagine 0.098 0.186 5.5 0.983 17.“0 (Asn) 0.103 0.0758 5 6 0.993“ 15.85 0.103 0.0759 5.5 0.9931 15.61 m.w. 150.1“ 0.101 0.131 6 1 0.988 1“.86 0.100 0.186 7.0 0.983 11.78 0.10“ 0.0381 7.1 0.9970 23.05 11“ TABLE 11 (cont'd.). Volume Ostwald Solution Solution Fraction Hydration Amino Solubility Molarity pH Hater Number Acid L M v H tw Leucine 0.103 0.15“ 6.“0 0.983 “.06 (Leu) 0.1009 0.15“0 6.9 0.98“3 11.65 0.10“0 0.1150 6.7 0.9881 3.35 m.w. 131.18 0.1057 0.0369 6.95 0.9965 -1.00 0.102“ 0.0750 6.8 0.9922 19.3 0.107 0.0751 6.67 0.992 -12.85 0.0989 0.0750 6.7 0.9920 “3.5 Phenylalanine 0.101 0.1“5 5.90 0.982 11.13 (Phe) 0.103 0.0686 6.51 0.991 15.57 0.1029 0.1“50 6.“ 0.9818 “.22 m.w. 165.19 0.1051 0.0351 6.85 0.9963 7.55 0.1052 0.0690 6.“ 0.9913 -0.92 0.1031 0.1070 6.“ 0.9870 7.“3 Sodium Chloride 0.088 0.500 0.991 17.80 (NaCl) 0.082 0.750 . 0.987 15.75 0.075 1.000 0.981 15.13 m.w. 58.““ 0.055 2.001 0.961 12.23 0.039 3.001 0.9“0 10.55 0.023 “.000 0.917 9.69 0.018 5.251 0.888 7.57 0.098 0.250 0.996 15.72 0.063 1.500 0.972 13.93 0.090 0.500 0.991 15.71 115 TABLE 12. Hydration numbers of the amino acids. Amino Acid This Work Reference Reference Reference (77) (78) (79) ALA 7.9 1 0.6 3.3 1 0.3 3.5 1 0.2 3.35 ARG 0.2 1 0.5 ------- “.1 1 0.5 ---- ASN 15.1 1 0.9 6.1 1 0.3 """" "" CYS 6.8 1 0.3 ------- “.3 1 0 2 ---- GLN 13.3 1 1.0 13.3 1 0.“ ------- ---- GLY 8.5 1 0.6 8.2 1 0.3 3.“ 1 0.3 3.“0 HIS 7.“ 1 0.5 35.6 1 2.8 “.6 1 0.8 “.62 HYP “.5 1 0.“ 16.2 1 0.8 ------- ---- ILE ------- 2“.0 1 0.“ ------- ---- LEU ------- 8.3 1 0.“ “.3 1 0.2 “.53 LYS 6.1 1.1 20.1 1 0.5 ------- ---- MET 8.1 1.0 21.1 1 0.8 “.5 1 0.5 “.53 PHE -------------- “.6 1 0.2 “.“8 PRO 2.0 1 0.2 29.9 1 1.6 3.3 1 0.3 3.26 SER 11.3 1 0.6 “.2 1 0.3 ------- “.0“ THR 7.6 1 0.5 -------------- “.06 VAL 8.0 1 0.7 “.9 1 0.“ 3.8 1 0 2 “.22 SUCROSE 3.9 1 o 7 25.8a ------- 3.u2 NaCl 16.2 1 0.5 ------- 20.2b ---- a Reference (81);b Reference (59) 116 The difficulty with calculating a hydration number for these five amino acids stems from their low solubility in water. Consider equation (66): H 55.3“60 _ L(M) I T[Vt ] o (66) w 0.1060 Since the experiments with LEU, PHE, TRP, and TYR all involved dilute solutions, with M S 0.15“, and the coefficient which multiplies the square bracket is large. For M << 1, both terms in the square brackets in equation (66) are within a few percent of unity. Thus, a 11 change in either v or L(M) results in a large change in H. The uncertainty in tw our values of L(M) are of the order of 1%. This uncertainty is probably the source of the large variation in calculated H values for these five amino acids. He must also mention a possible problem with the H values calculated for cysteine. In each of the runs, we observed a white precipitate forming during the course of the experiment. The precipitate is probably cystine, which forms from aqueous cysteine solutions in the presence of air. We estimate the mass of the precipitate to be about 0.56 g, which is the amount of cysteine which can form in the presence (1) v(2) 8 8 volume of V1 ~ 5 cm3). Cystine is very insoluble in water; a liter of water will dissolve only 0.112 g of cystine at 25°C. Thus, nearly all of 130 cm3 of air (V - 70 cm3, - 55 cm3, and the effective the cystine formed precipitates out of solution. We did not correct the measured solubilities for cysteine for changes in volume, molarity, or in vtw’ which occurred as a result of this effect. 117 The solubility data can be presented in several ways. Figure 30 is a plot of L versus M. Figure 31 is a plot of L versus Vt". In both graphs, the data for each amino acid can be approximated by a straight line which passes through the point for pure water. The linear dependence of L on both the solution molarity and the volume fraction of water suggests that the hydration number for each amino acid is a constant, i.e., each additional amino acid molecule decreases the volume of water available to disolve Xe by a constant amount. The assumption of constant H is borne out if we consider Figure 32, which is a graph of L/L, versus v w - HM/55.3“6 . The ordinate can t be recognized as the volume fraction of free water as measured by Xe solubility. The abscissa is the volume fraction of free water in a solution of molarity M and volume fraction of water v assuming a tw’ constant mean hydration number H. The data, for the most part, closely follow the line L/Lo - v w - fin/55.3116 . Deviations tend to be at high t solution molarity, where we might expect, as in NaCl solutions, a deviation from a constant H. What properties of amino acids determine the hydration number? Figure 33 is a plot of the mean hydration number as a function of the molecular weight of the amino acid, which is a rough measure of the size of the amino acid molecule. From this plot, we can see that there is no obvious correlation between H and the size of the molecule. The data are also broken down according to the type of side chain, whether it is 7 When viewed in this way, some nonpolar, polar, or positively charged. observations do suggest themselves. If we first consider the amino acids with nonpolar side groups, we see that the majority have hydration numbers in the range from 7 to 118 0°“ I I I I I 4:. + ALA 0.10 r- A Ana 3 A D LYS or! n ‘ - H . - + I; 0 09 E] 3 0 A30 ‘3 ommu- a + 4. U '3 omn - 3 C3 0.06 - LYS E 0.05 I I L I I mo me go L5 26 as Solution Molarity 0°“ F I I I I r I ‘\\b 4- ax >‘ 0.10 r- . A 99° 4.: or. \o‘ x U HYP i: MN 4 a + '3‘ x 3 ammu- ‘\‘~- H o n U) - . 0 00 F- -*L Q :3 ' u .A. Fan In 3 mm 4.) 8 0.07 1- ELY ...,,“ . LA 0.“ L I I I I I I (L0 as 14: L5 2m 21: 1L0 15 Solution Molarity Figure 30. L vs. M for Xe in aqueous amino acid solutions. 119 0.11 >5 4‘ 0.10 «4 r4 «4 .D a 000’ .1... C3 00 13 0.00 r4 (0 I! 13 CD 0.07 0.05 I I I I I. I I 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.8 Solution Molarity °-1°7 I I 1* I I I I -F VAL > 0.10 A HIS _, 4a ‘ E] GLN I. «4 _k.\ F4 . «4 0.10 2 “\.. + '- D 9\ :3 E] E] 4 ' r4 C3 .. - .— 0, 0.101 t! E] n - p4 VIL 4; :1 H18’ 4- U) I, I I I EL1 I I I 055* .. 0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40 Solution Molarity Figure 30. (cont'd.). 120 0. 107 r I I I I >5 .— 4; 0.105 At ..4 a A H «4 + + n A 0.10 + L - 3. o \\ 00 £1 MEJ’ A; g 0.101- 4. M51» -. + + ..I m A ILE ILE ASN In 0 0.09 A _. El 4- 0.09 L I g I I 0.00 0.04 0.00 0.12 0.10 0.20 0.24 Solution Molarity 0409 I I I. I I I f 5, 0.10%- + - 4.3 4-0 H «4 0.10 n :J H o (D 0.10 E m 0.101 I! 4.! In D 0.0 + —I 0.0 I? L I L I I I I 0.00 0.02 0.04 0.00 0.00 0.09 0.12 0.14 0.1! Solution Molarity Figure 30. (cont'd.). 121 Solution Molarity Figure 30. (cont'd.). ‘ ° F o —I f; 0 mci " «4 Sucrose .4 o - «4 3 o 4 H a . - ‘D 0 .... T In 5 ° '1 m J 0 ° + o «- ° '. Ostwald Solubility Ostwald Solubility Figure 31. L vs. v 122 Water tw for Xe in aqueous amino acid solutions. °4‘ I I I I I 00‘0 ~ - A’ an A A a 009 + - us a a 009 + _ + 4- MJ A. an 0.07 an :1 our - a 0.0. I I . l I I 0.09 0.90 0.92 0.94 0.95 0.90 1.00 Volume Fraction Water ” 0.00 I I [H 0.10 e/4 .. 0 P 04%; 009 a - ° A .0»: - 007 - mm mm 009 an - 0&5 1 00 07 0a 09 L0 Volume Fraction 123 0.11 0.10 '- 0.07 Ostwald Solubility III I“ 3 06+ 333 I L I I b 0.90 0.92 0.04 0.00 0.00 1.00 Volume Fraction Water, Ostwald Solubility .° 3 1‘ L I I I use 050 099 030 :00 Volume Fraction Water Figure 31. (cont'd.). 12“ 0.10 _I > «u A A El ’2 0.10 .2. :‘I A d o!" ‘9 mm + 3 0.9011- A =- v -1 o a) A 1:! + A + E. 0.09 _ A ++ a co + 0w 5", °-°‘7" am A m " m ‘ ASN U ASN O 0.095 + - 04m: ' l ' ' 0.975 0.990 0.955 0.990 0.995 . 1.000 Volume Fraction Water ' + .3-0Am5 + 0H r4 AL ‘5 4-0 '3 0¢m4 + - '3 A A m + t: 0.102 " H + 1.50 g + A we ‘3 _ o 0.900 O I + L I I 0'069900 0.995 0.990 0.995 1.000 Volume Fraction Water Figure 31. (cont'd.). Ostwald Solubility 0.11 0.10 0.00 0.00 0.07 0.06 0.06 0.04 0.03 0.02 0.01 125 9+ I NaCl Sucrose J I 4. -F I 0.5 0.6 0.7 0.0 0.9 Volume Fraction Water Figure 31. (cont'd.). 126 LIMI/Io o 5 I I I I 0.5 0.0 0.7 0.0 0.9 1.0 _ v,W - FIN/55 . 345 1.0 , , l 1 l I LIMI/Lo 0.3 I I I I I I 0.3 00‘ 005 °C. 0.7 00' 0.9 ‘00 Figure 32. L/Lo vs. vtw - HM/55.3“6 for amino acid solutions. 127 o ..I \ 3 ..I ' 0.5 0.5 0.7 0.5 0.9 1. V'w " HM/55.346 1.00 0.90 - o ..I 2 . 0.90 - 3:3 ..I 0.94 — 0 92 I I I ' 0.92 0.94 . 0.95 0.99 1.00 Figure 32. (cont'd.). 128 L (M1 /L0 I I I I 0‘11.“ 0.90 0.99 0.94 0.95 0.95 1.00 v,W - FIN/55 . 345 1.00 L (M) /L° Figure 32. (cont'd.). I I I 0.94 0.95 0.95 1.00 Hydration Number 129 ‘3 I I I I I I I I I 15 - A - 1‘ F'- I— 12 4 u C A _ 10 "' ‘5 0 " ‘ ‘- 7 - + A «— B - 4- «4 5 __ + 1 charge _‘ A polar . A g " 0 nonpolar '- 2 - 0 -( 1 - - o .. 4-.. _1 L I I I I I I I I I 70 60 00 100 1 10 120 130 140 150 160 170 160 Figure 33. Molecular Weight H vs. molecular weight. 130 9. Glycine, which has a molecular weight of 75.07, has a hydrogen atom as its side group, and although it is included in the polar group, it might also be included in the nonpolar group. GLY has an H of 8.5, which is in the proper range. The amino acids with polar side groups tend to have higher H values. The exceptions are THR, CYS, and HYP. Hydroxyproline will be discussed below. Cysteine has the difficulty discussed above, namely a large uncertainty in H due to the cystine precipitate. Threonine is the third anomalous amino acid in this group. We would expect, based on the similarity of structure of serine and threonine, that these two amino acids would have nearly identical hydration numbers. The large difference is puzzling. The amino acids with positively charged side groups may have slightly lower hydration numbers than the nonpolar group. It is interesting to note that ABC and LYS display a large difference in H, despite a strong similarity in structure. Arginine, with H - 0.2 1 0.5. probably has some affinity for Xe. Proline differs from other amino acids in that the nitrogen atom in the amino acid head group is also bound to the side chain, forming a ring structure. Since it is bound, the amino group is no longer ionized in solution. The low hydration number for PRO, H - 2.0 1 0.2, could be due to either a Xe affinity, or to the loss of the dipolar charge of the amino acid head group. The slightly higher value of H - “.5 1 0.“ for hydroxyproline, which is a proline molecule with an additional hydroxyl (OH) group, could be due to the addition of the polar OH group. Figure 3“ shows H versus pH, with the same breakdown according to R group as in Figure 33. There is no obvious correlation between H and 131 ‘5 I I I I I I I 16 - A d 14 b A + 1 charge " 13 " A polar - C. .— a) if .. A 0 nonpolar : I: C: 6 " Age’fib’ -F .— O 7 r A .- ..u-g + u 5 '- - m — - L 5 A ‘U 4 f- _ 2 " 0 _ 1 f- 4—1 0 I- + '- _1 I I I I I I I 4 5 5 7 a 9 10 11 12 Solution pH Figure 3“. H vs. solution pH. 132 pH, although there may be some decrease in H with increasing pH for PH>7. Eucken and Hertzberg59 have measured the solubility of the rare gases in a variety of alkali/halide salt solutions. The calculated hydration number of NaCl is H - 20.2, which agrees favorably with our value of 16.2 1 0.5. Goto80 reports a hydration value ofli-:22 for NaCl, calculated from effective volumes of electrolytes in aqueous solution. There are several other methods which may be used to calculate hydration numbers. These methods include ultrasonic interferometry, near infra-red spectrophotometry, and analysis of molal volume and adiabatic compressibility data. 77 Hollenberg and Ifft used the spectrophotometry method to measure hydration numbers of amino acids in solution. Hollenberg and Hall also used this method to measure hydration numbers of sugars. Some of their values are listed in Table 12. Also included in the table are 78 hydration numbers from the work of Millero et. a1. and of Goto and 79 The former analysed apparent molal volume and compressibility Isemura. data, and the latter used ultrasonic interferometry. Our values are included for comparison. The data of Hollenberg and Ifft are highly variable, even for similar molecules. For example, LEU and ILE have hydration numbers of 8.3 1 0.“ and 2“.0 1 0.“, respectively. This, along with other 82 considerations, probably indicate that we should discard their results. 133 What can we conclude about our hydration numbers? Why do our values differ so markedly from those of Millero_e_t_a_l. and Goto and Isemura? Our method of determining hydration numbers is quite different from that of either of these groups. Our values of H are actuallyaa measure of many different effects. We assume that the excluded volume effect is the only process which causes a change in solubility. This is certainly a naive model, since we are working with a three component system, and interactions between the components are undoubtedly very complex. At the least, there is probably a second contribution to the hydration values from Xe/amino acid interactions, such as is hinted at by the hydration number of arginine. The difference between our values, and those of Millero e_t §_1_. and Goto and Isemura, may be due to differences in the strength of the water/amino acid "bond". If water were tightly bound only to the polar head group of an amino acid, and not to the rest of the molecule, it is possible that only this tightly bound water would affect such bulk properties as the partial molal volume, adiabatic compressibility, and sound velocity. Although our model of amino acid hydration is not altogether well defined, it does explain the linear dependence of L on solution molarity. It also shows that there may be an attractive interaction between Xe and arginine in aqueous solution. There are several avenues of research to pursue here, also. First, a buffer might be added to the amino acid solutions to control the pH. This would be closer to physiological conditions in which the pH of the cell is controlled. A disadvantage to this approach is the addition of a fourth component to an already complex system. 13“ Second, experiments could be done with polyamino acids in aqueous solutions. The advantage here is that the contribution to hydration from the polarity of the amino acid head group is eliminated, since these groups are bound to each other in the formation of the polypeptide bonds. These experiments would be very expensive to undertake at the present time, but as improvements are made in the production process, the cost of polyamino acids should decrease. Third, studies might be done on the solubility of Xe in solutions «of the amino acids in the normal alkanes and alkanols, or in olive oil, which is a biological, nonpolar solvent. The liquid environment in these solvents is similar to that in the interior of a cell membrane. LIST OF REFERENCES 10. 11. 12. 13. 1“. 15. 16. 135 LIST OF REFERENCES G. L. Pollack, Rev. Mod. Phys. 36, 7“8 (196“). M. L. Klein and J. A. Venables, eds., Rare Gas Solids, vol. 1 and 2, (Academic Press, London, 1976). G. A. Cook, ed., Argon, Helium, and the Rare Gases (Interscience, New York, 1961). G. L. Pollack, U. S. Nuclear Regulatory Commission, Accession No. 8008190073, September, 1979; Accession No. 800“160039. March 1980. 'M. J. Halsey, R. A. Millar, and J. A. Sutton, eds., Molecular lwechanisms in General Anesthesia (Churchill Livingston, Edinburgh, 197“). I). E. Metzler, Biochemistry the Chemical Reactions of Living Cells (Academic Press, New York, 1977): L. Stryer, Biochemistry (W. H. Freeman and Company, San Francisco, 1975); A. L. Lehninger, Biochemistry (Worth Publishers, New York, 1972). J. Timmermans, Physico-Chemical Constants of Pure Organic Compounds, Volume 1, (Elsener Publishing Co., New York, 1950). R. C. Wilhoit and B. J. Zwolinski, J. Phys. Chem. Ref. Dat.,_2, Supplement 1, (1973). R. S. Berry, S. A. Rice, and J. Ross, Physical Chemistry (John Wiley and Sons, New York, 1980). CL Boato, G. Scoles, and M. E. Vallauri, Nuovo Cimento_l“, 735 (1959). H. C. Heber and H. P. Meissner, Thermodynamics for Chemical Engineers (John Wiley and Sons, New York, 1959). M. Tribus,‘Thermostatics and Thermodynamics (Van Nostrand, New Jersey, 1961). R. H. Davies, A. G. Duncan, G. Saville, and L. A. K. Staveley, T. Farad. Soc. 63, 855 (1967). J. C. G. Calado and L. A. K. Staveley, T. Farad. Soc._§]. 289 (1971). T. D. O'Sullivan and N. 0. Smith, J. Phys. Chem. 1“, 1“60 (1970). 17. 18. 19. 20. 21. 22. 23. 2“. 25. 26. 27. 28. 29. 30'. 31 . 32. 33. 3“. 35. 36. 37. 38. 39. “0. “1. A. 136 Ben-Naim, Water and Aqueous Solutions (Plenum Press, New York, 197“). G. J. Yevick and J. K. Percus, Phys. Rev. 101, 1186 (1956). J. K. Percus and G. J. Yevick, Phys. Rev. 191, 1192 (1956). J. K. Percus and G. J. Yevick, Nuovo Cimento 5, 65 (1957). J. K. Percus and G. J. Yevick, Nuovo Cimento 5, 1057 (1957). J. K. Percus and G. J. Yevick, Phys. Rev. 119, 1 (1958). E. Thiele, J. Chem. Phys. 32, “7“ (1963). M. S. Wertheim, Phys. Rev. Let. 19, 321 (1963). J. L. Lebowitz, Phys. Rev. A.l§§v 895 (196“). 11. Reiss, H. L. Frisch, and J. L. Lebowitz, J. Chem. Phys. 31, 369 (1959). H. Reiss, H. L. Frisch, E. Helfand, and J. L. Lebowitz, J. Chem. Phys. 32, 119 (1960). E. Helfand, H. Reiss, H. L. Frisch, and J. L. Lebowitz, J. Chem. Phys. §_3_. 1379 (1960). H. H. Uhlig, J. Phys. Chem. “1, 1215 (1937). D. Eley, T. Farad. Soc. ;§, 1281, 1021 (1938). Reiss, Advances in Chem. Phys. 9, 1 (196“). . Reiss and S. W. Mayer, J. Chem. Phys. 3“, 2001 (1961). . A. Pierotti, J. Phys. Chem. 61, 18“0 (1963). A. Pierotti, J. Phys. Chem. 62, 281 (1965). Wilhelm and R. Battino, J. Chem. Phys. 52, “012 (1971). J. Yosim and B. B. Owens, J. Chem. Phys. 32, 2222 (1963). J. Yosim, J. Chem. Phys. “9, 3069 (196“). C. Longuet-Higgins and B. Widom, Mol. Phys. 8, 5“9 (196“). S. Snider and T. M. Herrington, J. Chem. Phys. “1, 22“8 (1967). G. Kirkwood and F. P. Buff, J. Chem. Phys. 12. 77“ (1951). I. Prigogine, The Molecular Theory of Solutions (Northlkflland Publishing Company, Amsterdam, 1957). “2. 113. 11:. us. “6. 117. us. “9. 50. 51. 52. 53. 5“. 55. 56. 57. 58. 59. 60. 61. 62. 63. 6“. 65. 137 K. Watanabe and H. C. Anderson, J. Phys. Chem. 99, 132 (1986). A. Michels, T. Wassenaar, and P. Louwerse, Physica.g9, 99 (195“). A. Ben-Naim, J. Phys. Chem. 99, 792 (1978). A. Ben-Naim and Y. Marcus, J. Chem. Phys. 99, ““38 (198“). M. C. A. Donkersloot, J. Soln. Chem. 9, 293 (1979). K. J. Patil, J. Soln. Chem. 19, 315 (1981). K. Kojima, T. Kato, and H. Nomura, J. Soln. Chem. 19, 151 (198“). H. Schneider in Solute-Solvent Interactions,.J. F. Coetzee and C. D. Ritchie eds., (Marcel Dekker, New York, 1969). D. D. Van Slyke, J. Biol. Chem. 99, 3“7 (1917). D. D. Van Slyke and J. M. Neill, J. Biol. Chem. 91, 523 (192“). S. Y. Yeh and R. E. Peterson, J. Pharm. Sci. 99, “53 (1963). S. Y. Yeh and R. E. Peterson, J. Appl. Physiol. 39, 10“1 (1965). P. K. Weathersby and L. D. Homer, Undersea Biomed. Res. 7. 277 (1980). " J. Setschenow, Z. Physik. Chem. (Leipzig) 9, 117 (1889). S. K. Shoor and K. E. Gubbins, J. Phys. Chem. 19, “98 (1969). W. L. Masterton and T. P. Lee, J. Phys. Chem. 19, 1776 (1970). W. L. Masterton, J. Soln. Chem. 9, 523 (1975). A. Eucken and G. Hertzberg, Z. Physik. Chem. 195, 1 (1950). C. M. Lederer, J. M. Hollander, and I. Perlman, Table of Isotopes, sixth edition, (thn Wiley & Sons, New York, 1967). D. D. HOppes and F. J. Schima, eds., NBS Special Publication 626, "Nuclear Data for the Efficiency Calibration of Germanium Spectrometer Systems", (NBS, January 1982). R. T. Morrison and R. N. Boyd, Organic Chemistry (Allyn and Bacon, Boston, 1975). R. R. Dreisbach, Physical Properties of Chemical Compounds. II. (American Chemical Society, Washington, D.C., 1959). M. Windholz, ed., The Merck Index (Merck & Co., New Jersey, 1976). K. Siegbahn, ed., Alpha-, Beta-, and Gamma~ray Spectroscopy, volume 1, (North Holland, Amsterdam, 1965). 66. 67. 68. 69. 70. 71. 72. 73. 7“. 75. 76. 77. 78. 79. 80. 81. 82. 138 11. G. Cuming and C. J. Anson, Mathematics and Statistics for Technologists (Chemical Publishing Co., New York, 1967). G. L. Pollack, J. Chem. Phys. 19, 5875 (1981). G. L. Pollack and J. F. Himm, J. Chem. Phys. 11, 3221 (1982). G. L. Pollack, J. F. Himm, and J. J. Enyeart, J. Chem. Phys. 91_, 3239 (198“). G. L. Pollack and J. F. Himm, J. Chem. Phys. 99, “56 (1986). J. Makranczy, K. Megyery—Balog, L. Rusz, and L. Patyi, Hung. J. Ind. Chem. 9, 269 (1976). H. L. Clever, R. Battino, J. H. Saylor, and P. M. Gross, J. Phys. Chem. 91, 1078 (1957). H. L. Clever, J. Phys. Chem. 93. 375 (1958). V. G. Komarenko and V. G. Manzhelii, Ukr. Phys. J. 19, 273 (1968). M. H. Abraham, J. Am. Chem. Soc. 199, 2085 (1982). T. PL.1;illey, in Chemistry and Biochemistry of the Amino Acids, G. C. Barrett, ed., (Chapman and Hall, London, 1985). J. I... Hollenberg and J. B. Ifft, J. Phys. Chem. 99, 1938 (1982). F. J. Millero, A. LO Surdo, and C. Shin, J. Phys. Chem._9_2, 78“ (1978). S. Goto and T. Isemura, Bull. Chem. Soc. Jpn. 91, 1697 (196“). S.Goto, Bull. Chem. Soc. Jpn. 91, 1685 (196“). J. L. Hollenberg and D. 0. Hall, J. Phys. Chem. 91, 695 (1983). J.Jayne, J. Phys. Chem. 91, 527 (1983). U S "'11111111111111111111111111111111111111“