R? .‘o .u‘ _;.“f;¥i .JA' r}? ) "t v ." 4%? ‘: ‘52-» M‘ w‘? H 1;.“- at .P‘ . .. 3-,” -| 9' f-‘t‘flf'. -\ PK‘ «33.;- A; ‘ . r. f -_ r IIIJ 1‘ 1 :4 . w. . v" ,., L .. v‘ ..L Agtfff ‘ ‘1‘ 05., ‘ ‘ ‘ v“ Ky ' \' - :1“ ”AN.“ _ V 35...}, u“, u" ' 7‘ {FUJHJ'V vac" ”’4 " "‘V‘V' )~_ «1' ' 'l" ‘a n ‘ l ‘V ..;;;. a 1 ' .XF. , ,4»: 1 ‘fi fif‘gfit‘: - .‘ 2’, 1-9 1' “'41:?1 ‘ . _ ‘ . ‘ 57.x" ' .. w 2. 6:1 7 ‘7 ‘ $9.3 ‘I war ~==‘--'- “ w? 1:: I?“ ;/ ‘t 1.: .f ”Ht” This is to certify that the dissertation entitled REDUCED TRANSPORT OF ORGANIC POLLUTANTS IN SOIL MODIFIED WITH HEXADECYLTRIMETHYLAMMONIUM presented by Ines Toro-Suarez has been accepted towards fulfillment of the requirements for Ph.D. degree in Crop and Soil Sciences Environmental Toxicology Date%WéL¥/iia€~ MSU is an Affirmative Action/Equal Opportunity Inst/lunar! 0-12771 ill/ll!ill/IllllllllllllIll/llll/l/ll/llIll/Illllll 39040 F. LIBRARY Michigan atate University PLACE IN RETURN BOX to romovo thio checkout from your rocord. TO AVOID FINES roturn on or botoro doto duo. DATE DUE DATE DUE DATE DUE MSUlo oAnAflirmotivoActiorVEwolO Orippom ityl notitwon manna-m REDUCED TRANSPORT OF ORGANIC POLLUTANTS IN SOIL MODIFIED WITH HEXADECYLTRIMEI‘HYLAMMONIUM By Ines Taro-Suarez A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Crop and Soil Sciences Institute for Environmental Toxicology 1994 ABSTRACT REDUCED TRANSPORT OF ORGANIC POLLUTANTS IN SOIL MODIFIED WITH HEXADECYLTRIMEI'HYLAMMONIUM By Ines Tom-Suarez Groundwater use for drinking water in the U.S. justifies efforts to prevent or remediate groundwater contamination. This laboratory is developing a remediation technique that couples increased sorption of pollutants with microbial degradation. The overall objective of this research was to evaluate solute sorption during transport in Oshtemo Bt2 horizon soil exchanged with hexadecyltrimethylammonium (HDTMA). In the first part of this study, mechanisms of interaction between this soil and pentafluorobenzoic acid (PFBA) were evaluated. Anion exchange of I’FBA with cationic HDTMA, formed by HDTMA C16 alkyl moieties interacting by London dispersion forces, was demonstrated using 36Cl' and anionic PFBA in a buffer solution (pi-I 5.4). A PFBA mass balance demonstrated ion-pair formation between PFBA dissolved in water and desorbed HDTMA. Adsorption at the soil/water interface, adsorption at the HDTMA/water interface and partitioning between water and the HDTMA phase were the mechanisms considered responsible for retention of PFBA by I-IDTMA remaining in the soil after column equilibration. Adsorption at the HDTMA/water interface was the major contributor to retention. In the second part of the study, sorption coefficients (K) from water for naphthalene, benzene, and TCE were extrapolated from the predicted linear relationship between log K values obtained in water-methanol mixture from soil column breakthrough curves (BTCs), and the corresponding methanol fraction (Rao et al., 1985). These values were compared to sorption coefficients from sorption isotherms. The two methods did not coincide, and the relationship between log K obtained from batch isotherms and the corresponding methanol fraction was not linear. Benzene sorption coefficients from water obtained from BTCs and batch isotherms in soil with the same organic carbon content were different. Benzene also did not follow the linear relationship between log K and methanol fractions (0 to 0.40) predicted by Rao’s model. These results could be explained by methanol/HDTMA interaction and the need to redefine the phase ratio for I-IDTMA-modified soils. Evidence for methanol/HDTMA interaction was indicated by a strong affinity of methanol for the octane/water interface. ACKNOWLEDGMENTS I would like to thank my advisor, Dr. Stephen A. Boyd for financial support during my graduate work and for give me the facilities to growth in a new area. He arranged a cooperation with the University of Florida, where I had the privilege to learn the secrets of soil columns personally from Dr. P.S.C. Rao. I had the opportunity to work with Dr. Maya Gugeshashvili, with her expertise I was able to study interfacial adsorption. I also had, for some time, the technieal assistant of Christopher Gallo. With Surash, Maya and Chris I am personally thankful. Each one of the members of my committee (Dr. S. A. Boyd, Dr. SJ. Anderson, Dr. V. McGuffin, Dr. J. Tiedje and Dr. M. Zabik) has contributed in a very specific way to my professional growth. For this. I am very thankful. I specially enjoyed the enlighting discussions with Dr. McGuffin. I would like to acknowledge the editorial work, at different points in the writing process, of Linda Schimmelpfennig. Dr. Sharon Anderson, Mary Ann Bruns, Drs Judith and Ray Frankmman and Dr. Donald Christenson. Special thanks to Mary Ann for her consistent help. The close cooperation between Dr. Tiedje and Dr. Boyd’s groups was always intellectually and socially rewarding, as well as, a source of friendship and moral support. I I would like to recognize the help in multiple occasions of Darlene Johnson, iv Jerry Wardwell, Linda Salemka, Dr. Dave Harris, Dr. Boyd Ellis, Dr. Delbert Mokma, and Dr. Lee Jacobs. TABLE OF CONTENTS . Page LIST OF TABLES ..................................... vm LIST OF FIGURES .................................... x CHAPTER 1 ........................................ 1 CHAPTER 2 ........................................ 22 2 1 Introduction ................................ 22 2.2 Materials and Methods ......................... 25 2.3 Results and Discussion ......................... 29 2.4 Summary .................................. 36 References ................................. 39 CHAPTER 3 ........................................ 40 3.1 Introduction . . . . . . . ........................ 40 3.2 Materials and Methods ......................... 45 3.3 Results and Discussion ......................... 51 References ................................. 67 CHAPTER 4 ........................................ 69 4.1 Introduction ................................ 69 4.2 Theory ................................... 72 4.3 Materials and Methods ......................... 76 4.4 vii Table 2-1 Table 2-2 Table 2-3 Table 3—1 Table 3—2 Table 3-3 Table 3—4 Table 4-1 Table 4-2 Table 4-3 Table 44 LIST OF TABLES Page Properties of Oshtemo Bt horizon soil. ................ 26 Anion exchange capacity (ABC) and amount of HDTMA at saturation coverage in an HDTMA modified soil. ......... 30 Comparison of CEC values obtained by different methods. 34 Properties of Oshtemo Bt horizon soil. ................ 46 Surface tension measurement at the air-water and octanol-water mterface. .................................. 59 Adsorption parameters of anionic PFBA in the octane-water system at 22°C. .............................. 63 Comparison between soil surface areas measured using N2 (BET) and estiamted using PFBA. .................. 65 Properties of Oshtemo Bt horizon soil ................ 77 General characteristics of the columns ................ 80 Retardation factor (R) and sorption coefficient (Kw) obtained from BTC. Kw obtained from batdh sorption isotherm. ..... 87 Properties of the native soil and HDTMA modified Oshtemo soil and of soil columns prepared with these soils. viii Table 4-5 Table 4-6 Summary of methanol adsorption parameters at the octane—water SummaryoforganiccarbonpresentinthesoilafterBTC experiments from an initial organic carbon content in the soil ix Figure 1-1. Figure 1—2. Figure 1-3. Figure 1-4. Figure 2-1. Figure 2-2. Figure 2-3. Figure 2-4. Figure 3-1. Figure 3-2. Figure 3-3. Figure 3-4. LIST OF FIGURES Page Response and schematic representation of useful parameters for pulse (A) and step (B) input functions. ................ 4 Danckwers’ F-diagrams of fluid flow thorugh columns. ..... 7 Brenner solution for equation 1. .................... ll Graphical representaiton of R calculation using method 4 in Nkedi-Kizza, et al., 1987. ....................... 15 Effect of dispersion in the breakthrough curve shape. ....... 23 35cr - Anion exchange on HDTMA-modified soil. ........ 31 PFBA adsorption isotherms on HDTMA-modified Oshtemo soil. ................................... 32 HDTMA mass balance in HDTMA-modified soil and PFBA water system ................................... 37 Schematic diagram of miscible-displacement system. ....... 47 Breakthrough curve of PFBA in Oshtemo Bt horizon soil. . . . . 52 Breakthrough curve of PFBA/buffer in HDTMA-modified Oshtemo Bt horizon soil ......................... 53 Change in interfacial adsorption (y) as a function of concentration. ............................. 60 Figure 3-5. Figure 4-1. Figure 4-2. Figure 4-3. Figure 4-4. Figure 4-5. Figure 4-6. Figure 4-7. Figure 4-8. Figure 4-9. Figure 4-10. Figure 4-11. Excess surface concentration (1‘) vs PFBA concentration. Schematic diagram of miscible-displacement system. Log of naphthalene sorption coefficients (K) at different fraction of methanol (fc). Effect of HDTMA soil modification in benzene breakthrough curve (BTC) in water. Effect of HDTMA soil modification on TCE breakthrough curve (BTC) in water. Benzene sorption coefficient (K) obtained from BTCs and sorption isotherms in different fraction of methanol (fc). TCE sorption coefficient (K) obtained from BTCs and sorption isotherms at different proportions of methanol (fc). Change of interfacial tension 7 in the octane/water interface as a function of methanol concentration. HDTMA desorption during miscible displacement experiments. . Benzene sorption coefficient (K) estiamted form BTCs, obtained in seven columns and from sorption isotherms obtained in soil from these columns. Results of fitting benzene breakthrough curve data to a linear and quadratic equation. xi 62 ........................... 100 ..................... 102 CHAPTERI Introduction . It is estimated that more than 40% of the U.S. population utilizes groundwater asasourceofdrinkingwaterandthat25% ofallfreshwaterusedintheU.S.is groundwater. Thus, it is important to understand and predict pollutant transport from the soil surface through the unsaturated and on saturated zones to groundwater (Bouchardet et al. , 1989). The geologic materials through which these compounds move are very complex porous media. This porous medium has two major components: solid particles and free space between the particles and within their pores. Organic matter and minerals (present in clay, silt, and sand-sized particles) in different proportions constitute the soil solids. A dissolved organic compound entering the soil will move with the bulk flow or be retained due to interaction with the organic matter and/or the mineral components of the soil. It may also diffuse into the small pores of soil particles. Therefore, the main parameters to be considered in the study of the transport of organic compounds through natural porous media are: water flux, molecular diffusion of the compounds, organic carbon content of soil or aquifer material, and water solubility of the compounds. The behavior of a compound in a soil or aquifer material can not be accurately described unless all of these parameters are considered simultaneously. ’ To study the interaction of these variables at the laboratory level, soil and environmental scientists have normally used what they call miscible-displacement experiments. Miscible—displacement experiments are performed by passing a solution of the solute to be studied through a column filled with the material to be characterized (i.e. , soil, sand, aquifer material, or sediments). The change in concentration is followed 1 2 at the outlet of the column by collecting fractions for further analysis or by continuously feeding the outlet of the column through a concentration-sensitive detector. This method was the original form of chromatography and is also known as frontal chromatography. The response from frontal chromatography is transformed to obtain the effluent breakthrough curves (BTCs). Tiselius (1940) introduced the theory of frontal chromatography which served as a base for the further development of chromatography theories. These theories have been used to explain the behavior of compounds of environmental concern flowing through natural porous materials. The overall objective of this research has been to evaluate the transport of organic contaminants in soil columns. The soil contained in the columns has been previously modified by substituting the native inorganic exchangeable cations with a quaternary ammonium cation, hexadecyltrimethylammonium (HDTMA). This has been shown in batch equilibration experiments to substantially increase the sorptive properties of the soil for nonionic organic contaminants (NOCs) (Boyd, et al., 1988, 1991; Lee et al., 1989). In the present study, miscible-displacement or frontal chromatography has been tools used to study transport of organic contaminants and the processes that influence their mobility. Thus, the purpose of this chapter is to indicate how environmental scientists have quantitatively related the chromatographic retention and dispersion parameters of the BTCs to physical and chemical characteristics of the system. Components of a Miscible-displacernent Experiment The behavior of water flowing through a soil or aquifer material is of the same nature as that of any fluid flowing through a bed of stationary solid particles. For this reason, it is obvious that the same mathematical equations as models applied in general to study the movement of solutes through porous materials can be used to study the flow of organic contaminants through soil and aquifer materials. Thus, soil scientist have taken advantage of and improved the theories advanced by investigators in other related 3 fields, such as the theories developed for chromatography and porous media (Nielson and Biggar, 1962). In this chapter the different aspects needed to study the behavior of organic contaminants flowing through geological materials will be considered in a systematic order. Selection of the Input Function. In any chromatographic experiment, the type of chromatogram obtained (the response) depends on the shape of the sample input profile (the input function). Reilley et a1. (1962) studied the chromatographic response corresponding to different input functions. Figure 1 shows the response profiles for the two most common types of input functions, the impulse and the step functions. In A, a concentration of the sample is introduced instantaneously in the column (impulse function). In B, the concentration of sample entering the column increases rapidly from zero to a finite value and remains constant at that value (step function). The step injection mode is referenced to as frontal chromatography. Chemical engineers and soil and environmental scientists have called this form of sample injection miscible-displacement, due to the fact that the sample is mixed in the porous medium at the same time that it moves through the column (Nielsen and Biggar, 1961). Selection of the input function depends on the objectives of the analysis and the column diameter. During miscible-displacement experiments, the solid particles restrain, deflect, and disperse the flow. The form of the flow pattern is determined by the geometry of the interstitial pore space. The front of the zone advancing through the column is spread (dispersed) due to different velocities in the flow stream. The internal structure of the porous medium also influences the diffusion rates. Therefore, it is very important to know as many details as possible of the flow pathway (Giddings, 1965). Two types of dispersion can occur. Radial dispersion occurs when a sample is injected in the impulse mode. Radial dispersion is much slower than axial dispersion at R! 8'0"“ l . [rel :l . dh—c- non-“.1 . dDuOv—I. ——-0 1r ~§ r. I '1 1; I I I I I I I I I I Figure 1. Response and schematic representation of useful parameters for pulse (A) and step (B) input functions. From Riley, et al. (1962). 5 normal flow rates. Thus, the band of solute will travel through the column contained within a central core of the packing. Depending on the column diameter and the particle sizediameters, thecentralcorecanbesocontainedthatinmanycasesthesolutenever reaches the column wall. Knox (1977) quantified this phenomenon with chromatographic columns and called it the infinite diameter effect because the column behaves as it has a very large diameter. Axial dispersion occurs more rapidly close to the walls of the column. Therefore, in chromatography it is considered advantageous for the sample to avoid contact with the column walls because it decreases spreading in the signal. However, this is not the case when part of the objectives of the research is to study dispersion effects due to column packing, as is the case with soil and aquifer materials. This is one of the reason why during this type of research, it is more precise to use the step mode of injection. In this injection mode the sample volume is big enough that the solution will contact almost all the parts of the packing material. The other reason to use the step mode instead of the impulse mode during the characterization of the dynamic behavior of soils, is that due to the broad range of particle size present in these materials, wider bore columns must be used to minimize walls effect. With the impulse injection mode, the response or peak height of a compound decreases as its retention increases . This is a dispersion effect called the column dilution factor (Knox, 1977) and is greater in larger columns. In the Step mode. the compound solution is introduced in the column until the response is equal to the input concentration. In studies with soil columns, the compound is sorbed in the column until it reaches equilibrium. The response of the solute concentration (Co) is established first outside of the column. Thus, equilibrium is assumed to be reached when the response of the compound eluting from the column is equal to the Co response. Therefore. to be able to use this equilibrium criterion, it is necessary to use the step injection mode. It is also possible to use the step method to study the effects in sorption caused by varying solute concentration (adsorption isotherms) (Reilley et al. ,1962). 6 Now that the reasons for selecting the step function to perform the frontal chromatographic or miscible displacement experiments with soil columns has been established, the construction of the ETC will be considered. Breakthrough Curves Danckwerts (1953) studied the problems presented by steady- flow in different systems and showed how its behavior can be investigated and quantitatively specified. He defined the breakthrough curve (or 'F- diagram“) as the plot of F(t) vs vt/V, where F(t) is the fraction of concentration of the compound at the outlet of the column as a function of time and W V is the ratio between the total volume that has flowed through the column at time t and the volumetric water content. Lapidus and Amudson (1952) used a similar approach to represent the change in concentration at the outlet of the column but they used the total volume of fluid instead of the volumetric water content. Nielsen and Biggar (1962) adapted Danckwerts’ procedure in their studies with soil columns, adopting the notation of C/Co for F(t) and referred to the volumetric ratio as pore volumes. This is still the most common notation used today. Lapidus and Admundson (1952) were the first to demonstrate, with chromatographic columns, the influence of flow velocity and diffusion in the shape of the BTCs. Danckwerts also showed that his F- diagrams (BTCs) revealed a good deal of information about the behavior of the fluid flowing through the column. In Figure 2, taken from his paper, the shapes of different BTCs are presented. As he indicated, they represent different behavior of the fluid as influenced by the column packing. Figure 2a corresponds to the ideal behavior of a fluid through a column where elements of fluids which enter the column at the same time will move through it with constant and equal velocity on parallel paths and will leave the column at the same time. Figure 2b is indicating deviation from ideal behavior due to some mixing of the fluid in the column. Figure 2c corresponds to the shape of the BTC when perfect mixing of the fluid is o b e . d i V F e r a - r o r '. F "W —"' ' (a) Pink.» flow; (it) l'bttott flow with (c) Complete training; (cl) Dead voter. tantra longitudinal tttixhtm Figure 2. Danckwers’ F—diagrams of fluid flow through columns. (Danckwerts, P.F. , 1953). 8 occurring in the column, whereas Figure 2d represents the diagram obtained where there is dead water present in the system (stagnant water trapped in pores). Nielson and Biggar (1962) indicated that if any solute flowing through the column is retained within the column by any chemical or physical process, the breakthrough curve will move to the right. This can be concluded by simple inspection of the compound BTC. Thus, the breakthrough curve is descriptive of the relative time taken for the displacing fluid to flow through the medium or for the solutes in the displacing fluid to come in chemical equilibrium with the medium. Hydrodynamic Dispersion The ideal behavior of a fluid flowing through a column is known as piston flow (Figure 2a). However, in reality the fluids are Newtonian in nature. There will always be some spreading in the column profile (Danckwerts, 1953). This longitudinal spreading is called hydrodynamic dispersion and is due to a continuous exchange of velocities in the flow stream. The configuration of the porous medium is very complex. It is constituted of cavities, pores of different sizes, channels, and micropores inside the particles, all randomly distributed in the medium. Therefore, different points in the flow stream have different velocities; this has been called the microflow pattern (Giddings, 1963). Spreading depends on how rapidly these velocities can be exchanged. Two mechanisms determine how fast flow exchange occurs for a compound traveling through a column without interaction with solid material (tracer). The first one is molecular diffusion, a process by which a molecule can move from one flow path to another. The second mechanism has been called eddy diffusion, which occurs when the compound follows the flow path that randomly changes velocities during its travel through the column. In chromatography, hydrodynamic dispersion has been studied using a theory known as the coupling theory of eddy diffusion (Giddings and Robison, 1962). In porous media, it has been well established that if the macroscopic flow is 9 accepted as one-dimensional, the longitudinal dispersion may be described by the one- dimensional diffusion equation in which the diffusion coefficient is replaced by the coefficient of dispersion (D). E = DE - ”E (1) Of 512 OZ Where, C(x,t) is concentration in solution (mollcm3) t is time (s) x is distance from the inlet (cm) D is the dispersion coefficient (cm2/s) U is the linear velocity (cm/s) Also the length of the bed is L and the initial solute concentration is Co. When this equation is solved for a non-retained compound, we have the simple case of miscible displacement. In this case, the only effect that influences the movement of the tracer is the flow phenomenon. Any compound flowing through the system, whether it is retained or not, will sense this effect. Brenner (1962) gave solutions to equation (1) in a general form and in tables, where the dimensionless exit concentration is prwented as a function of dimensionless time (T = Ut/L) for different Peclet numbers (by analogy with the convective heat) which defined as P = 111141) where D is the axial dispersion coefficient. There are several techniques described in the literature to determine P from the BTC of a tracer. The two most common will be considered here. Hashimoto et al. (1964) determine P by fitting the BTC data of a tracer to the Brenner solutions by a least square method. Then, the best fit is obtained by interpolating in graphs such as the one presented in Figure 3. Rose and Passioura (1971) developed a method to determine P in which the data are transformed in a semilogarithmic form. The data fall on a straight 10 line, and P is determined from the slope. P is related to the effective diffusion coefficient, according to Brenners’ defintion, through the equation: p . fl (2) 4D This equation can be solved with just the experimental data. So far we have indicated only the effects of dispersion on the shape of the breakthrough curve for non-retained compounds. The shape of the BTC is also influenced by the sorption isotherm and the mass-transfer kinetics for compounds interacting with the media. Sorption Isotherms The use of frontal chromatography to measure the sorption isotherm of solutes on solid surfaces is the oldest and, for many years, the most common application of this technique (Parcher, 1978). When a compound has a non—linear sorption isotherm, retention will change with concentration. For this reason, only frontal chromatography can be used for this process; To determine the adsorption isotherm, the amount of solute adsorbed in the solid material is determined at different concentrations (Gluckauf, 1945). As with any BTC, the effluent concentration from the column is monitored until it reaches a constant value equal to the inlet concentration. This indicates the saturation of the bed by the compound under study and the end of the sorption process. This method was recently employed by Thibaud et al. (1992) to investigate the adsorption equilibria of several volatile organic contaminants on soil. They determined the sorption capacity of each compound on the soil, at saturation, and compared to the adsorption capacity corresponding to a monolayer. They found that all the compounds studied had a BET type II adsorption isotherm and that the contaminants cover the whole surface uniformly. They also calculated, from the linear part of the adsorption isotherm, the BRENNER SOLUTION [Nrrnrcuctivc and Conservative Salutes] ~-u—i==co (Lotti-in; stun.) Figure 3. Brenner solution for equation 1 (Rose and Passioura, 1971) 12 molar heats of adsorption. When a compound has a non-linear sorption isotherm, the boundaries of its BTC are either diffuse or self-sharpening depending of the curvature of the isotherm. Also, if sorption takes place but the amount retained changes with concentration, our computational procedure can produce only an empirical description of column behavior which may possibly be useful but certainly will have no immediate theoretical interpretation. In consequence, a condition that must be met to be able to use the convective-dispersive equation is that the compound under study should have a linear sorption isotherm in the medium. That is, the ratio of the amount of solute in the mobile phase to that in the stationary phase, the sorption coefficient (Kd), should be constant and independent of solute concentration. Reilley, et al. (1962) concluded that if a compound has a linear isotherm in a partition chromatographic column, the retention time (t, ) remains constant over the range of concentrations used. More importantly, the shape of response in each step, sorption and desorption, are symmetrical. Thus, if the sorption isotherm is linear, other factors will cause changes in the shape (symmetry) of the BTC. Retardation Factor (R) If'linear reversible equilibrium sorption is assumed, sorbed and aqueous solute concentrations may be related using the partition or distribution coefficient, Kd, such that: s = K; (3) where S is the concentration of solute sorbed by the soil (mg/kg) and C is the concentration of solute in solution at equilibrium (mg/L). Hashimoto et al. (1964) solved the diffusion equation for a column of finite length under the assumption of a linear 13 sorption isotherm. They defined a dimensionless factor that they called the retardation factor (R) expressed as: R=1+%xy ‘ (4) Where p is the bulk density and 0 the volumetric water content of the column. Thus, the difusion equation can be rewritten: 2 6C_D6C 5C (5) __ ———U— D! 5x2 bx The retardation factor R is a quantitative index of a chemical compound’s mobility, in that its value is equal to the ratio of the positions of the sorbed and non-sorbed solute in the soil column. For a non-sorbed solute, the value of the sorption coefficienth in the equation is equal to zero: hence, R = 1. For sorbed solutes, R is greater than one since the value of Kd is larger than zero. A compound with a large value of R has reduced mobility in soils. Since there was a lack of a standard method to determine the retardation factor from BTCs, Nkedi-Kizza et. al. (1987) decided to evaluate four different methods. The compounds selected were diuron and atrazine, two. pesticides with non-linear sorption isotherms. In the first method, the values of R was calculated from sorption isotherms in mixed solvents using the equation I pK'Co 0H) 0 R" = l + 0 98% . All chemicals were used without further purification. HDTMA-modified soil. An Oshtemo Bt2 horizon (coarse-loamy, mixed, mesic Typic Hapludalt) was used to prepare the hexadecyltrimethylammonium (HDTMA) modified soil. The soil (500 g) was mixed with 1000 ml water while an aqueous solution of HDTMA (8.2 g in 3000 ml) was added in an amount equal to the cation exchange capacity of the soil (Table 1). The suspension was stirred overnight, then transferred to centrifuge bottles and centrifuged at 2,603 g for 20 min. The supernatant was removed using a gentle vacuum suction, and the bottles were refilled with water (225 ml). This procedure was repeated three times to remove any excess HDTMA. After the washing process, the soil was air dried. 1“c-rrnTMA-rnodified soil. The basic procedure used to prepare the HDTMA-modified soil, described above, was also used to prepare the l4C-HDTMA modified soil. In addition to the HDTMA 46 required to satisfy the soil CBC, 0.1 ml of l4C-I-IDTMA solution was dissolved in the aqueous solution added to the soil suspesion containing 100 g of soil. Table 1. Properties of Oshtemo Bt horizon soil Particle Size (96) Sand 89 Silt 5 Clay 6 CBC (mmol/Kg) 45 pH 5.8 flame Carbon Content 0.1 Miscible Displacement Experiments Miscible displacement experiments were used to determine the hydrodynamic dispersion of the column (Peclet number) containing nonmodified soil and to determine the retention of PFBA in the HDTMA-modified soil. A schematic diagram of the miscible displacement system is presented in Figure l. The system consists of two glass bottles, one containing the solute (the sample) and the second the elution solvent. Two single piston pumps (Model 302, Gilson Medical Electronics Inc.) and (Model 510, Waters Ass.) are used in the constant pressure mode for sample and solvent delivery. The pumps are connected to the column through a four- way solvent selection valve (Model 5020, Reodyne). The inlet of the column is also connected to the four way solvent selection valve. The outlet of the column is connected to a flowthrough ultra-violet/visible variable-wavelength detector (Holochrome, Gilson). The column isconnected to the system with low dead-volume Teflon fittings and tubing. The various connections are kept as short as possible to minimize extra-column 47 589$ EoEoomEflee—afie he :5..ch cum—cocoa A «...—«E assoc :20 25 A 05:. t|/|\ anyhow—mo (DU) _ 2...; 232.5 _ e528 Ill/p I _ ..ow not 58% N mine c258 58F 48 contribution to dispersion. Changes in concentration as sensed by the detector were registered on a chart recorder (ABB model SE 120). Three columns were used during these experiments. One was filled with non- modified Osthetemo Bt horizon soil and two with the HDTMA-modified Oshtemo soil, one with l4C-HDTMA and the other nonlabeled. The column filled with the nonlabeled modified soil was used to obtain the BTC for PFBA/buffer solution, and the labeled one to obtain a soil equilibrated under similar conditions as those used when the BTC was performed. The column used was a borosilicate glass preparative column (Candies, Chromaflex) with an internal diameter of 2.5 cm and a length of 5 cm. Bed supports consisted of woven Teflon (FE) diffusion mesh and a Teflon (TEE) filter membrane (extra-fine pore-2.5 um) mesh and a 0.45-ttm nylon filter membrane. The fraction of air-dried HDTMA-modified soil that passed a 1-mm diameter sieve was used to fill the columns. The columns were dry-packed using the "rotate, bounce and tap method” (Knox 1976) widely used for packing materials with particle sizes higher than 40 am. The dry material was packed to obtain a uniform bulk density. For pro-equilibration, the columns were connected, in standing vertical position, to a piston pump (Model 302. Gilson Medical Electronics, Inc.). The columns were equilibrated with pure water until steady-state water-saturation conditions were established. A buffer solution containing a mixture of 0.5 N potasium phosphate monobasic and 0.05 N NaOH (pH 5.4) filtered through a 0.45-ttm Millipore system and degassed with helium was used to prepare a PFBA solution (50 mg L’l) and to elute the column. After determining the inlet solution UV deflection in the recorder (Co-response) at 235 nm, the BTC of PFBA was measured by displacing the sample solution through the column, at a linear velocity of 90 cm h'l. The sample was displaced until the detector 49 response reached Co. Then the eluent was switched to plain buffer solution. The column was eluted until the effluent UV response was equal to the original base line. HDTMA desorption During the column experiment with the 14C-l-IDTMA-modified soil, eluent was collected and its volume determined. The l4C-activity in these solutions was determined in triplicate by LSC. OctanollWater Partition Coefficient Equal volumes of buffer solution (0.5 N KI-12PO4 and 0.05 N NaOH) and octanol (Fluka, Ronkoma, N.Y.) were allowed to equilibrate for four days. PFBA was dissolved in the octanol-saturated buffer solution to obtain a concentration of 500 pg m1“ 1 . Three 20-ml glass vials were filled with 15 ml of the PFBA/buffer solution and allowed to equilibrate in an incubation chamber (New Brunswick Sci., Edison, N .J .) at 25 i 1°C. Then, 5 ml of water-saturated octanol was added to each of the vials, which were sealed with teflon-lined septa. The vials were stirred for 48 hrs. in an incubation chamber (25 j: 1°C) at approximately 200 rpm. A sample of the octanol layer was removed from each vial using a Paster pipette, and a sample from the water layer was drawn off with a long stainless steel needle connected to a glass syringe to avoid disturbing the octanol layer. The PFBA concentrations in the two layers were measured using a diode array UV/VIS spectrophotometer (Hewlett Packard, model 8452A). The K0W was calculated as the ratio of the concentration of PFBA in octanol to the concentration of PFBA in the buffer solution. Also, Na+ and K+ were extracted back in water from 1 ml aliquot of the octanol phase. The Na+ and K+ concentration in the water extract as well as in the water phase, were determined in a plasma emission spectrophotometer (Spectra Span, model VB). ‘ 50 Interfacial Adsorption of PFBA Equal volumes of octane and 0.5 N KHZPO4 buffer (pH 5.4) were equilibrated for 48 hrs. A series of PFBA solutions were prepared in buffer saturated with octane solution and equilibrated for four days with an equal volume of octane saturated with buffer. Using these solutions, the interfacial tension at the octanol/water (buffer) and air/water (buffer) interfaces was measured by the drop method. The interfacial tension measurement apparatus consisted of a micrometer, syringe, capillary tube and container. A buffer solution containing PFBA was delivered to the end of one capillary tube from a syringe, whose plunger is operated by a micrometer. The tip of the capillary must be ground smooth so that the end is sharp, regular, free from any nicks, and perpendicular to the tube. The syringe was previously calibrated to determine the volume of liquid per unit of micrometer scale. As the liquid is delivered from the capillary to the immiscible second phase, a drop forms and eventually breaks away, and the volume of the drop is determined by reading the micrometer. Absorption spectra of the bulk solutions were measured by a Hewlett Packard (Model 8452A) Diode Array spectrophotometer to determine PFBA concentrations at equilibrium. This value was used to correct PFBA initial concentration. T h e densities of water and organic solutions were measured using a pycnometer with a 2 ml volume. The interfacial tension 7 was obtained from the equation Y = “d1 'd2)8f r 51 where (11 and d2 are the densities of the immiscible liquids studied, V is the volume of the drop, g is the gravitational acceleration, f a correction factor (from Hrakins-Brown table), and r the radius of the tip of the capillary. The change of interfacial tension as a function of change in solute concentration (PFBA) was used to construct the first type adsorption isotherm, described in detail in the Results and Discussion section. 3.3 Results and D'scussion Oshta-o Bt horizon soil PFBA interaction To evaluate possible interactions between PFBA and the native soil, a BTC of PFBA dissolved in 0.01 N CaC12 was obtained in a column filled with the unmodified soil. Figure 2 shows the BTC obtained with this column, where the relative concentration at the column outlet is plotted against the number of column pore volumes. The calculated retardation factor for PFBA was 1.0, which indicates that the compound is moving with the front of the solvent and is not being retained by the soil. The shape of the BTC is sigmoidal and symmetric indicating that nonequilibrium, as would be indicated by tailing is not present in the system. PFBA interactions with the HDTMA-modified soil In contrast to the observed behavior of PFBA in columns filled with the unmodified Oshetmo Bt horizon soil, we observed retention of PFBA in the HDTMA- modified soil. This is shown clearly in Fig. 3 which presents the BTC for PFBA (Co 50 mg/ L or 0.24 mM in 1012904 buffer at pH 5.4) in the HDTMA-modified soil column. The calculated retardation factor was equal to 57.4 pore volumes. It is interesting to note that it took 26.2 pore volumes for the outlet concentration to be above the detector signal detection limit. Some tailing is observed in the complete BTC (sorption and desorption) indicating the presence of some mass transfer nonequilibrium. 52 0.9- 0.8‘ 0.7- 0.6- 0.5- C/Co 0.4- ‘ 0.3- 0.2- 0.1- Ivent: 0.01 N CaCI2 0 . 1' i 5 - Pore volumes fi-l Figure 2. Breakthrough Curve of PFBA in Oshtemo Bt horizon soil 53 0.9- I - 0.8- 0.7d C/Co O ‘l' I 0.3- - ' 0.2- 0.1-‘ - $ 50 too 150 200 250 Pore volumes Figure 3. Breakthrough Curve of PFBA/Buffer in HDTMA-modifed Oshtemo Bt horizon soil 54 Previous studies on sorption characteristics of NOCs in HDTMA-modified clays and soils conducted under equilibrium conditions have shown that the organic matter normalized partition coefficients (Kom) of NOCs sorbed by HDTMA-modified subsoils were 10to30timesgreaterthanthoseobservedonnativesurfacesoils. Intheformer case, the organic sorption phase is composed almost entirely by HDTMA whereas in the latter case it is natural soil organic matter. The interacting hydrocarbon moieties of HDTMA exchanged in soil clays forms a more effective partition phase for NOCs titan natural soil organic matter, which contains an abundance of polar oxygen functional groups such as hydroxyls and carboxyls. In comparison, the hydrocarbon tails of HDTMA form a phase that closely resembles a reverse-phase chromatographic material (e.g. C18) where the long hydrocarbon tails covalently bonded to the -SiOH functional groups of silica gel (Snyder and Kirkland, 1978) form an organic phase that is compositionally similar to the HDTMA-derived phase. Thus, the theory and behavior of organic acids in reverse-phase chromatography may be useful in understanding the interaction of anionic PFBA with the HDTMA-modified soil. Since the development of partition chromatography (Martin and Synge, 1941) organic acids have been studied in non-polar chromatographic materials using a simplified model, where the compounds are assumed to partition between the polar mobile phase and the nonpolar stationary phase covering the solid support. However, these studies have largely ignored the role of interfacial adsorption in retention. Ion-pair formation between hydrophobic acids and quaternary alkylammonium ions was used by Wahlund (1975) as a method to separate hydrophobic acids in reversed-phase chromatographic columns. Partitioning of ion pairs between water and the nonpolar phase, previously covered with l-pentanol, was the assumed mechanism. Adsorption of the PFBA at the HDTMA/water interface is considered here as a plausible mechanism to account for the strong retention of PFBA. Martin (1961) was the first to suggest that solute adsorption at the gas-liquid interface can contribute to 55 retention in gas-liquid chromatography. Adsorption of polar or ionic compounds at the interfacebetweenthenonpolarstationaryphaseandthegasphase,inreversephasegas chromatographic systems, was later demonstrated by Bereekin (1978). For reverse phase high pressure liquid chromatography, however, no report was found where interfacial adsorption has been experimentally evaluated. To account quantitatively for the strong retention of PFBA by HDTMA-modified soil, we attempted to consider separately the contribution of PFBA (or PFBA ion pairs) dissolution into the HDTMA phase and contribution of interfacial adsorption. To evaluate the magnitude of these two processes, the total retention of PFBA in the HDTMA-modified soil was compared to (a) PFBA partitioning between octanol and water, and (b) the interfacial adsorption of PFBA in octane-water mixtures. Model solvent selection Under the conditions in which the ETC experiment was initially conducted, the compound was completely ionized. This was because the pH of the system was 4.8 which is 3 units greater than the pKa of PFBA (pKa = 1.48). Thus, in the model binary solvent systems evaluated (octanol/water and octane/water) pH was equal to or above 4.8, so that PFBA in the aqueous phase was completely ionized. During the interfacial adsorption experiments phosphate buffer (pH 5.4) was used to keep PFBA completely dissociated. At pH values where acidic compounds are nearly completely ionized (pH > pKa + 2), the formation of ion pairs has been shown to increase with ionic strength (Kaiser and Valdmanis 1981; Westall et al. 1985; Jafvert et al. 1990; Lee et al. 1990). Although the high ionic strength of the buffer solution used in these experiment (a = 0.55) would favor the formation of ion-pairs, we attempted to minimize ion-pair formation by using monovalent cations in the buffer solution (0.5 N potassium phosphate nonobasic and 0.05 N NaOH). Role of Partitioning in PFBA 56 The octanol-water partition coefficient (Kow) of PFBA was determined to be 0.04. To establish if ionic pentafluorobenzoate or neutral ion pair had been transferred to octanol, the presence of cations in the buffer solution, and in the octanol fiaction, was determined. The presence of Na+ in the octanol phase (Table 1) indicates that PFBA was probably transferred to the octanol phase as an ion pair with Na+ from the buffer solution. The concentration of Na+ in octanol (36.5 mg L'l) was higher than the PFBA concentration (20.5 mg 1:1) indicating that excess Na+ had partitioned into the octanol fiaction. Evidence for both ion pairs and free ions in water-saturated octanol was also reported Westall et al. (1990) in a study of the distribution of several strong electrolytes between water and octanol. It is noteworthy that K+ , present in the buffer solution at a concentration 10 times higher than Na+, was not detected in the octanol phase. Westall et al. (1990) also found that Na+ more readily partitioned into octanol, consistent with the size of the hydrated ions and the Bohr model. Since PFBA is a hard ligand, it is expected to complex more strongly with a harder metal ion (Na+) that has lost its hydration shell in a medium with a lower dielectric constant (octanol) than in water. Another important factor may be the size of the ion: Na+ is smaller than K+ , and it will be more readily complexed with a ligand having a small bite (aperture) like PFBA. At this point, we can only speculate about the reasons why Na+ is more favorably transfered in the octanol phase. However, this finding is interesting enough to justify further research. Using the experimentally-determined Kow value, the maximum amount of PFBA retained by the soil as a result of ion-pair partitioning was calculated as an estimate of the contribution of this mechanism (PFBA partitioning into the HDTMA organic phase) to the overall retention of PFBA in the soil column. The distribution coefficient of PFBA in the soil is defined as: 57 = [pmlm 0" [PFBAJW where [PFBAJHDTMA is equal to the total number of pg of PFBA per unit volume of HDTMA (ml) and [PFBA1bufl-fl. is the number of pg of PFBA per ml of water contained in the column. If we assume that KOM for PFBA in the soil is equal to KOW (Jaynes and Boyd 1991) and we know the HDTMA content of the soil, the total number of mmols of PFBA retained by the soil in the column (soil loading capacity) can be obtained as follows: = [_|.tg PFBA/ml HDTMA] 0" [pg PFBA/ml water] The volume of HDTMA present in the column is estimated from the density (p) of hexadecane: ‘ HDTMA weigg p hexadecane Vol HDTMA = The HDTMA content of the soil (0.3 %) is estimated from the equilibrated solid organic carbon content (0.4%) minus the initial organic carbon content of the soil (0.1%). Knowing the mass of organic carbon in a mole of HDTMA (228 g), then the number of moles of HDTMA in 100 g of soil = 0.3 gC/228 gC = 0.0013 moles/ 100 g. Assuming that PFBA will dissolve only in the hexadecane tails, the weight of the sorptive phase in 100 g of soil will be the number of moles times hexadecane molecular weight (226.45 g), or 0.0013 moles times 226.45 g/mole = 0.299 g per 100 g of soil. The weight of hexadecane in the experimental column containing 45.42 g soil is 0.136 g. The volume of HDTMA in the soil column then is obtained by dividing the weight of HDTMA by 58 the density of hexadecane: 0.136 g/0.773 g/ml a: 0.176 ml. The denominator of the distribution coefficient is: = [Total pg PFBA - X] [pg PFBA/ml Wt] timer content (nil) Where X equals the mass of PFBA retained by the column or the loading capacity of the column, and total pg PFBA (in the column) is equal to the PFBA input concentration (52.2 pg/ml) times the volumetric water content (10.36 ml). Therefore: [(10.36 ml 1 52.2 Int!) - X] PF minute = [ns BA! r] 1036 ml Thus, assume Kow, determined experimentally as 0.04, is equal to Kom’ then: K _ 0.04 = [rt/0.176 ml] 0“ [(540.79 pg - xyrosomn Solving the equation for X: X = 0.366 pg or 1.73 x 10" mmoles These results indicate that the HDTMA-modified soil column could retain a maximum of 0.366 pg PFBA or 1.73 x 10’6 mmoles by partition interactions, if all of the above assumptions are true. Role of Interfacial Adsorption in PFBA Retention The most common way for studying adsorption of organic compounds at liquid/air or liquid/liquid interfaces is based on measurement of the interfacial tension. We used 59 the drop-weight method to evaluate the potential for adsorption of PFBA at the interface between water and the HDTMA-derived organic phase (referred to hereafter as the ”hexadecane-water interface”) of the modified soil. Model Selection. Octanol-water has been used since 1964 (Fujita et al.) as a model system to predict certain behavior of organic pollutants in the environment, as for example NOC partitioning into soil organic matter and the bioaccumulation is NOCs. We used the drop-weight method to evaluate the adsorption of PFBA at the octanol/water interface. Table 2 summarizes the PFBA adsorption data in the equilibrated octanol/water system and at the air/water interface. Table 2. Surface tension measurement at the air-water and octanol-water interface. System Surface tension (70) Surface tension (1) Surface pressure mN/m mN/m (70 - y) mN/m octanol/water 8.75 8.56 0. 19 (0.05 M PFBA) air/water 72.39 63.21 9.18 (plus 0.05 M PFBA) air/water 72.39 37.51 34.88 (saturat. with octanol) air/water 72. 39 37.5 1 34.88 (saturat. with octanol plus 0.05 M PFBA) Interfacial tension at the octanol/water interface '70 was equal to 8.75 mN/m. The addition of PFBA (0.05 M) did not substantially change the interfacial tension (7) of the .lb 3.5- I 0.5- U 0.606 0.608 0.01 0.612 0.014 Concentration (M) 0 0.602 0.604 Figure 4. Change in interfacial adsorption (7) as a function of concentration. 0L ”6 61 octanol/water system (10 - 7 = 0.19). Apparently the high surface activity of octanol obscures the surface activity of PFBA at the octanol/water interface due to the competition between two amphiphilic compounds. The surface activity of octanol, and the competition between octanol and PFBA, was clearly observable at the air/water interface. The surface tension at the air/water interface (70) is equal 72.39 mN/m. If 0.05 M PFBA was dissolved in the aqueous phase instead of octanol the decrease in surface tension (70 - 7) was only 9.18 mN/m. If octanol saturated water is used instead of pure water to determine the surface tension (7) at the air/water interface, the value decreases to 37.51 mN/M (10 - 'y = 34.88), due to the high surface activity of octanol. Thus, if PFBA is added to water saturated by octanol, the surface tension at the air/ water interface does not change significantly (Table 2). This precludes the use of octanol-water mixtures as a model to study PFBA adsorption at the interface between water and the HDTMA-derived organic phase. Thus, octane water was selected as the model system to study the interfacial adsorption of PFBA at the hexadecane/water interface of the HDTMA modified soil. Figure 4 represents. changes in octane/water interfacial tension as a result of increasing PFBA concentration in solution. As the graph shows, the octane/ water interfacial tension increases as a function of PFBA concentration indicating that PFBA is a surface-active compound. When the adsorption isotherm is plotted as I‘, where I‘ is the excess of particles of a substance adsorbed per unit of suface area of interface in relation to the concentration of the same substance present in the bulk solution (Gibbs, 1906) versus the PFBA concentration (Fig. 5), the maximal value of Gibbs surface excess (I‘m) is obtained. The value of I‘m corresponds to the coverage of the interface by a monolayer of PFBA. The minimal interfacial area per adsorbed PFBA molecule can be calculated from the equation: .10 I' x1 0 , mole/cm2 62 0.8 0.74 0.6- 0.5- 0.4- 0.34 0.2- 0.1- 0 0.601 0.602 0.603 0.604 0.605 0.606 0.607 0.608 Concentration (M) Figure 5. Excess surface concentration (1‘) vs PFBA concentration. 0.609 0.1 31 63 r - “Np" where N A is Avogadro number and A is the interfacial. area per adsorbed molecule. Thus, in the compact layer of the electric double layer at the octane/water interface All 8 (r-NA)-1 where Am is the minimal area per adsorbed PFBA molecule. This value, calculated from the I‘m obtained from figure 5, is given in Table 3. Table 3. Adsorption parameters of anionic PFBA in the octane-water system at 22°C. System I‘m 10'10 Am a -AG p M/cm2 Azlmol kJ/ mole PFBA (0.5M ' 0.716 231.84 0.455 23.73 0.63 KB P04 + 0.05 M Naéfl) - The PFBA adsorption parameters at the octane/ water interface can be calculated from the general form of the adsorption isotherm of amphiphilic compounds (Markin et al. , 1992): in - (p - 061'" a 0 _ exp(-2a0) = X expl-AG IRH p'fl-Q)’ A 64 where p is the number of adsorbed octane molecules substituted for by one molecule of PFBA, a is the attraction constant between the particles adsorbed, X is mole fraction of PFBA, AGO is the standard Gibbs free energy of adsorption at equilibrium, and 0 is the surface coverage. Using the software Mathematica (Wolfiam Research), the adsorption parameters p, a, and AG0 were calculated and are shown in Table 3. Time data are consistent with anionic PFBA adsorbed at the octane-water interface with the COO'-groups oriented into water and screened from the electric double layer by inorganic cations such as H+ , K+ , and Na+. The phenyl rings of PFBA are present in hydrophobic octane phase interacting due to Van der Waals forces; this is the source of the resulting attraction constant of 0.455. The interaction energy, 23.73 kJ per mole (5.68 Kcal/mol) corresponds to the sum of all Van der Waals forces between the adsorbed moieties. Estimation of Anionic PFBA Retention Due to Interfacial Adsorption To integrate the results obtained in the octane/water system with the results observed in the BTC, it has been assumed that, at the moment when PFBA started to elute from the column, a compact layer of PFBA has been formed at HDTMA/water interface .of the soil. Therefore, the minimum area A1 covered by a molecule of PFBA at the compact layer in the model octane-water system was used to calculate the surface area of the soil covered by the anionic PFBA, as shown below. The number of molecules retained by the soil before any PFBA elution occurs was calculated as follows: V [PFBA] = mass of PFBA adsorbed at the base line 65 where V (271.53 ml) is equal to the number of base line pore volumes before PFBA elution commenced multiplied by the volume of one pore volume. The input concentration of PFBA [PFBA] = 52.2 pg/ml. Therefore, 251.53 ml x 52.20 pg/ml = 14173.75 pg, or 0.067 mmols. Multiplying by Avogadros’s number gives the corresponding number of molecules: 6.02 x 10'23 molecules/mol x 0.067 mmol/1000 = 40.243 x 1010 Now, the number of molecules is multiplied by the area covered per molecule (Table 3) at the compact. layer: so, 4.024 x 1019 molecules. 231.84 A2/molecule = 9,330.1 x 1021 A2 To convert this value to square meter we divide by 1020: 9,330.1 x 1021 A211020 = 93.301 rn2 Thus, the maximum surface area coverage by adsorbed PFBA per gram of modified soil, obtained after dividing by the column soil weight (45.42 g), yields a value of 2.05 m2/g. In Table 3 the surface area of the soil calculated with the model value is compared to the value obtained using the BET method. Table 3. Comparison between soil surface areas measured using N2 (BET) and estimated using PFBA Method of Determination Surface Area (mzlg) BET 2.36 BTC base line and area cover/molecule 2.04 Interestingly, a very close agreement exists between the surface area calculated from the octane/water model and the N2 BET surface area of the modified soil. This would seem to be a reasonable result supporting the concept of interfacial adsorption of PFBA in the modified soil, realizing that the majority of surface area in this soil is due to clay mineral surfaces which are presumably covered with exchanged HDTMA. 66 Conder et al. (1969) established that the different contributions to gas- chromatographic retention (partition, adsorption at solid/liquid and liquid/liquid interfaces) are additive, and we have adapted this idea and applied to our system. The retardation factor (R) calculated from the BTC first moment and expressed column pore volumes for anionic PFBA in the HDTMA-modified soil column will be: Rtotal = (R)Ad.Soil/water + (Rldisol. + (”Ads Liquid/liquid Where the total retention Rtotal obtained from the BTC is equal to 57.4 pore volumes, the retention due to interaction of PFBA with the native soil (R) Ad. Soil/water is 1 pore volume, the number of pore volumes due to dissolution (mdisol is equal to 3.4, and the number of pore volumes due to interfacial adsorption in the HDTMA/water interface (R) Ads lquidlliquid is equal to 26.2 pore volumes. The sum of all terms in the right side is equal to 30.6, which is 26.8 pore volumes smaller than the Rtotal° The difference is approximately equal to the number of pore volumes due to interfacial adsorption. This could conceivably be accounted for by the formation of a second PFBA layer at the interface. At any rate, these results suggest that among the different contributors to retention of PFBA in the HDTMA modified soil, adsorption of PFBA at the hexadecane- water inbrface of the HDTMA-soil is the primary mechanism accounting for the strong retention of PFBA. 67 References Adamson, A.W. 1960. Interscience Publishers, Inc. New York, N.Y. Berezkin, V.G. 1978. J. Chromatogr. 159. 359. Boguslasvssgii, L. 1., A.N. Frumkin, and M. I. Gugeshashvili. 1976. Electrokhimiya. 12, Boyd, S.A., J-F. Lee, and M.M. Mortland. 1988a. Nature. 333, 345-347. Boyd S.A., M.M. Mortland, and GT. Chiou. 1988b. Soil Sci. Soc. Am. J. 52,652. Boyd, S.A.,W.F. Jaynes, and BS. Ross. 1991. pp. 181-200. In In R.A. Baker (ed), vol. 1, Lewis Publishers, Chelsea, MI. Bowman, R. S. 1984. Soil Sci. Soc. Am. J. 48, 987. Bowman, R. S. 1992. Ground Water, 30, 8. Brusseaué4M. g“, R. E. Jessup, and P. S. C. Rao. 1991. Environ. Sci. Technol. 26, 1 -14 . Burris, D. R. and C. P. Antworth, 1992. J. Contain. Hydrol. 10, 325-337. Conder, J. R., D. C. Locke, and J. H. Parnell. 1969. J. Phys. Chem. 73, 700-708. Damaskin, B. B. 1964. Doklady Acad. Nauk SSSR, 156, 128. Damaskin, B. B., O. Petrii, and V. Batrokov. 1971. Plenum press, New York- I.ondon. . Fujita, T., J. Iwasa, and C. Hansch. 1964. J. Am. Che. Soc. 86, 5175. Getmanskii, I. K., and L. I. Bavika. Assay Methods of Aqueous Surfactant Solutions (in Russian), 1965. Part 1, NIITEKhM, Mocow. Gibbs, J .W. 1906. Dover Reprint, 1931. Longmans and Green & Co., New York, N.Y. Gugeshashvili, M. 1., M. A. Manvelyan, and L. I. Boguslavskii. 1974. Electr'okhimiya, 10, 819. Jaynes, W.F. and S.A. Boyd. 1991. Soil Sci. Soc. Am. J. 55, 1991. L6e,J.F., J.R. Crum, and S.A. Boyd. 1989. Environ. Sci.Technol. 23,1365-1372. Lee, L. S., P.S.C. Rao, P. Nkedi-Kim, and JJ. Delfino. 1990. Envir. Sci. Tecnol. 24, 654. Lee, L. S., P. S. C. Rao, and M. L. Brusseau. 1991. Envir. Sci. Tecnol. 2, 722. Lyklema, J. 1991. Vol. I. Adacemic Press Inc. San Diego, Ca. 68 Markin, V. S., M. I. Gugeshasvili, A. G. Volkov, G. Munger, and R. Leblanc, 1992. J. Electrochem. Soc., 139, 455. Markin, V. S. and A. G. Volkov. 1989. Progr. Surface Sci., 30, 233. Martin, A. J. P. and R. L. Synge. 1941. Biochem. J. 35, 1358. Martin R.L. 1961. Anal. Chem. 33, 347. Me Bride, M.B. Surface chemistry of soil minerals. In Minerals in soil environments. 1989. SSSA Book Series, no. 1. Madison, WI. Mbellenberg, K., C. Leuenberger, and R. P. Schwarzenbach. 1984. Environ. Sci. Technol. 18, 625. Reilley, (29.8N.2, g3. P. Hildebrand, and J. W. Ashley, Jr. 1962. Anal. Chem. 34, 11 -1 l . Rittich, B. and M. Pirochtova. 1990. J. Chromatogr. 523, 227. Tomlingora, B., T. M. Jefferies, and C. M. Riley. 1978. J. Chromatogr. 159, 315- 5 . Wahlund, K-G. 1975. J. Chromatogr. 115, 411-422. CHAPTER 4 Directly Measured vs Predicted Sorption Coefficients During Transport for Nonionic Organic Compounds in a Hexadecyltrimethylammonium Modified S011. 4.1 Introduction Sorption and degradation (biological or chemical) of nonionic organic compounds (NOCs) cause decreased transport through soils, thereby reducing ground water contamination. Sorption of N OCs by soil and sediments in aqueous systems is controlled mainly by the organic carbon content of the sorbing material and by the water solubility of the NOC. With these facts in mind, this laboratory has been working to develop an integrated approach to remediate contaminated soils and aquifer materials. This approach consists of reducing transport of common ground water contaminants and then subsequently degrading them microbiologically (Nye et a1. , 1993). To reduce transport, the sorptive capability of sandy soils has been increased by exchanging the inorganic native cations of the soils with quaternary ammonium compounds (cationic surfactants) of the form (CH3)3NR where R is a large alkyl hydrocarbon. This method has been proven to increase the sorption of N OCs on surfactant treated B-horizon soils by over two orders of magnitude (Boyd et al. , 1988). Sorption studies with these modified soils indicate that the surfactant-derived sorptive phases are 10 to 30 times more effective than native soil organic matter as partition media for NOCs (Lee et al. , 1989). The two tools most commonly used to study sorption behavior of organic compounds in soils and sediments are batch sorption isotherms and breakthrough curves (BTC). With these methods, the thermodynamic distribution constant, or the sorption coefficient, K, of the compounds are determined. The sorption coefficient in the batch 69 70 method corresponds to the slope of the sorption isotherm constructed by plotting the concentration of the sorbed compound in soil versus its concentration in water contacting that soil at equilibrium. From the BTCs the retardation factor (R) of the compound, a parameter indicating the mobility of a solute eluted through a soil column, is obtained. K is determined from the R value by using the equation R=l+r£ (1) where p is the bulk density and 0 is the volumetric water content of the soil. The phase ratio, p/0, is the ratio between the two phases in the system, in this case the solid phase and the liquid phase, in which the compound flowing through the system will be distributed. The basic difference between the two methods is that one value is obtained under equilibrium conditions (batch method) and the other under dynamic conditions (BTCs). However, since K is an equilibrium parameter the two values should be the same. For natural soils, in most cases, these values have been shown to be the same (Nkedi-Kizza et al., 1987; Wood et al., 1990). The results obtained with laboratory soil columns are considered to provide a better understanding of the transport behavior of NOCs through soil or aquifer materials. Before this work, only batch sorption isotherms had been used to characterize the impact of chemical modification, with cationic surfactants, on the sorption capacity of soils. The first objective of this part of the study was to compare K values obtained during transport, i.e., from BTCs, with K values obtained from sorption isotherms. To obtain the ETC, a constant concentration of the compound must be fed into the column until equilibrium is obtained (step function injection mode). Based on the sorption coefficients obtained with the batch method, NOCs are expected to be highly retained in the soils and thus, their R values are expected to be high. This suggests an operational limitation to obtain the BTC directly from water, for volatile organic compounds, 71 because the elution volumes and hence elution times required are large. Nkedi-Kim et al. (1985) suggested for solutes with low water solubility and large sorption coefficients that sorption coefficients from water can be estimated fiom BTCs obtained from mixtures of water and organic solvents using a model developed by Rao et al. (1985). This model predicts a log-linear relationship between K and the fraction of organic solvent in water (fa. This relationship was developed by combining the Karickhoff (1984) equation for sorption of hydrophobic organic compounds (HOC) from water on soils and sediments with a solubility model developed by Amidon et al. (1974) and Yalkosky et al. (1975). A summary of the theory is presented in the next section. The theory was used by Nkedi-Kim et al. (1985) to predict the sorption of diuron, atrazine and anthracene from water in five different soils. They used different proportions of methanol-water and acetone-water. They found that, as expected, a log- linear relationship exits between the K values of the compounds and the corresponding fractions (fc) of methanol-water. However, when different proportions of acetone-water were used as solvent mixtures. a quadratic relationship appeared to fit the data better. The overall goal of this part of the study was to evaluate the retention of common ground water contaminants (benzene. naphthalene. and TCE) in soil columns containing quaternary ammonium cation modified soils. We compared the K values determined in columns filled with hexadecyltrimethylammonium (HDTMA) modified soil with the K values obtained in the same soil from batch equilibrium sorption isotherms. Because NOCs are strongly retained in the organically modified soils (Boyd et al., 1988; Lee of al. , 1989), the approach was to use the log-linear relationship between K and fc of water- methanol mixtures (Rao et al. , 1985) to estimate K values in pure water by extrapolation. Because the K values, for the compounds above, obtained by extrapolation using the model did not coincide with the K values calculated from the sorption isotherms and also because it was observed that the organic carbon content of the soil was reduced significantly during the BTC experiments. The values of K from effluent BTCs were obtained directly from pure water and compared to the extrapolated Ks. Also, K values 72 were obtained by batch sorption isotherms using soil removed from the column at the end of the BTC experiments. 4.2 Theory The model of Rao et al. ( 1985) was developed to describe quantitatively the sorption and transport of hydrophobic organic chemicals (HOC) dissolved in water and mixtures of water-organic solvents in sorbent materials such as soils and sediments. These sorbents are known to contain both mineral and organic constituents. However, based on Karickhoff (1984) findings, it was assumed that the surfaces in direct contact with the interstitial solution were predominantly hydrocarbonaceous in nature. The sorbate molecules were considered to be composed of hydrocarbon and polar moieties, with hydrocarbonaceous portions predominating. The theory considered sorbate-solvent and solvent-solvent interactions to be most important in describing HOC sorption on soils and sediments from single or mixed solvents. Sorption (sorbate-sorbent interaction) is assumed to be controlled by sorbate solubility in the solvent from which sorption occurs. Thus, the major theoretical assumption was that sorbate-solvent hydrophobic forces are responsible for sorption. Changes in sorbent physico-chemical properties due to solvent- sorbent interactions, as well as competition between solvent and sorbate for the sorbent surface, were neglected in the theory (Rao et al. 1985). Karickhoff (1984) showed that for soil and sediments, the logarithm of the HOC sorption coefficient normalized by the soil organic carbon content (PW) is equal to: AS In P a in X [—1] (T. T) 0 where a and 6 are constants, X‘" is the mole fraction solubility, ASf is entropy of fusion, Tm is the melting point (0K), T is temperature (0K) and R the gas constant (KJ/mol 0K). Karickhoff (1984) found the values of a and B to be 0.83 and 2.142, respectively, in soil and sediments. 73 Amidon et al. (1974) and Yalkowsky et al. (1975) modified the basic concept of solubility of Hildebrand et al. (1970), which was used to describe solvent-solute interactions in regular solutions, to study solubility in polar solvents. In an ideal solution all intermolecular forces are equal and there is no change in heat or volume during mixing. In a non-ideal solution, the deviation fiom ideal is known to arise as a result of the unequal intermolecular forces between solvent-solvent and solvent-solute molecules, and it is described by the activity coefficient (ac). Thus, the mole fiactional solute solubility (X2) is expressed as: dog X. do: xi“ + los (ac) 6’ The ac is considered to reflect the work required to remove the solute from its own environment (W22), plus the work required to create a cavity in the solvent large enough to contain the solute molecule (W 1 1), plus the work gained after the molecule is inserted into the cavity (W 12). Mathematically the log of ac can be expressed as: (W22 + W11 ' 2W12) "2"2 (4) 2.303” ”108 (a) = where V2 is partial molal volume, and 411 is the solvent volume fraction. For regular solutions, where the size and polarity of the solvent do not differ greatly, Hildebrand et al. (1970) approximated the term W12 by the geometric mean of W11 and W22. Thus, the equation for diluted solutions (4’1) becomes: V2 2.303RT log (ac) = - (wt? - W3): (5) where the square roots of the work terms are known as solubility parameters 61 and 32 of the solute and the solvent, respectively. When Yalkowsky et al. (1975) considered that for solutions in aqueous or polar solvents, the geometric mean approximation is not valid, they proposed a two dimensional analog of the equation of Hildebrand et al. ( 1970). Thus, instead of using the geometric mean approximation to obtain W12, they 74 used the concept that the work required to remove a solute molecule from the bulk phase is equal to the work of cohesion of the solute. This work is equal to the surface area of the solute (A) times the surface tension of the solute (11). The work required to create a cavity of area A in the solvent is considered to be equal to the work of cohesion of the solvent, i.e. , A72, where 72 is the solvent surface tension. Also, the work involved in the insertion of the solute molecule into the solvent corresponds to the work of adhesion, which is equal to A712, where 112 is the interfacial tension between solute and solvent. Thus, they arrived at the equation: '12‘2 (6) 2.303RT 4% (ac) = - Furthermore, considering that both interfacial tension and molecular surface area vary with structure, they adopted the approach of Langmuir (1911) of separating the contributions to interfacial tension of the nonpolar and polar portions of the molecule. Thus, equation 6 becomes: YlPAP +71?“ (7) 2.303” - 108 (a) = where 71b is the microscopic aliphatic hydrocarbon-solvent interfacial tension, and 71p is an analogous two-dimensional term dependent on the interaction between the solvent and the polar portion of the solute. Rao et al. (1975) adapted this basic concept of solubility. They expressed Yalkowsky mole fraction solubility for any crystalline solute in any pure solvent as: m,- g _ [tr/HSA) + (e'mn _ (ll—St)“ - 7) (s) kT RT ' where yj and ej correspond to ‘Ylh and 71p, respectively; HSA and PSA correspond to Ah and AP, respectively, according to Yalkosky et al. (1975) nomenclature; and k is the Boltzmann constant (kJ/K). They also used an extension of the cavity model developed by Yalkowsky et al. (1976) to predict HOC solubility in binary solvent 75 mixtures which is expressed as a linear combination of terms representing the pair-wise interactions of each solvent with each topographic component of the solute molecule: In X" = 11: X" +f‘[(Ay‘HSA) + (A e‘PSA)]IkT (9) where A 7 = (7w - 7° ); Ac 0 = (tsw - c); and the superscripts w, m and c correspond to water, mixed-solvent, and organic miscible solvent (cosolvent), respectively. Rao et al. ( 1985) extended the Karickhoff (1984) sorption equation to mixed solvents: AS 1n P alnX [ )(T. 7). B By substituting lnXm as defined by Yalkosky et al. (1976) in equation 9, they obtained the following equation: mp~=-m . , 11 x" + filmy HSA)” (Ae‘PSA)] _ (35.1](1'. _ 1) _ p ( ) Next, Rao et al. (1985) defined the term 0‘ = [(Ay‘HSA)k; (AE‘PSAH (11b) Combining Eq.(2) with Eq.(lla) and (11b) gives lnP" = lnP“'- ao‘ffc (12) The above equation expresses the log-linear relationship between the normalized sorption coefficient and the fraction of organic cosolvent. Rao et al. (1985) extended the concept of retardation factor to considered the effect of cosolvents on the transport of HOCs in soils: 76 p K" (13) R.=1+ where Rm is the solute retardation factor in a mixed solvent system, KIn is the non- normalized sorption coefficient and 0m is the liquid phase content (cm3 / cm3) They generalize Eq (11) for the non-normalized sorption coefficient and substituted in Eq (12) rearranged giving: log (R. - 1) .. rogue, — 1) - «07¢ (14) This equation predicts an exponential decrease of (Rm - 1) with increasing f0 and was the equation used in this research to obtain K in water by extrapolation. 4.3 Materials and Methods All solutions were prepared with reverse osmosis deionized water, purified further through a Milli-Q system (Milliport, Milford, MA). Benzene and TCE, both 99 % purity, were purchased from Aldrich Chem. Co. (Milwaukee, WI); naphthalene and hexadecyltrimethylammonium bromide, both 99 % purity, from Sigma Chem. Co. (St. Louis, MO); methanol, high purity, from Burdick & Jackson (Muskegon, MI); and octane from Fluka (Ronkoma, NY). l4C-HDTMA-Br, labelled on the terminal carbon of the hexadecyl chain, was obtained from Moravek Biochemicals, Inc. (La Brea, CA); it had a specific activity of 55 mCi mmol'1 and a radiochemical purity of > 98 %. All chemicals were used without further purification. HDTMA-modified soil. An Oshtemo Bt2 horizon (coarse-loamy, mixed, mesic Typic Hapludalt) was used to prepare the hexadecyltrimethylammonium (HDTMA) modified soil. The soil (500 g) was nlixed with 1000 ml water, after which an aqueous solution of HDTMA (8.2 g in 3000 ml) was added in an amount equal to the cation exchange capacity of the soil 77 (Table 1). The suspension was stirred overnight, then transferred to centrifuge bottles and centrifuged at 2,603 g for 40 min. The supernatant was removed using gentle vacuum suction, and the bottles were refilled with water (225 ml). This procedure was repeated three times to remove any excess HDTMA. After the washing process, the soil was air dried. Table 1. Properties of Oshtemo Bt horizon soil Particle Size (%) Sand 89 Silt 5 Clay 6 CBC (mmol/Kg) 45 pH 5.8 giant Carbon Content (%) 0.1 14C-HDT‘MA-modified soil. The basic procedure used to prepare the nonradioactive HDTMA-modified soil was also used to prepare the "enema modified soil. In addition to the HDTMA required to satisfy the soil sec. 0.8 ml of “(Z-HDTMA solution (3 mCi in 31 ml ethanol) was dissolved in the aqueous solution added to the soil suspension containing 800 g of soil. The resulting specific soil activity and organic carbon content were determined with 50 mg of finely ground soil samples. Triplicate samples of the soil were combusted at 1020°C in a Carbon Nitrogen -Mass spectrometer (CN-MS) analyzer (Roboprep, Europa Scientific, U.K.). The C02 released from the samples was split, a small portion was used in the mass spectrometer for carbon analysis, and the rest was collected. in liquid scintillation vials and analyzed on a Packard 1500 TriCarb Liquid Scintillation Analyzer (Packard Instrument Co., Downer’s Grove, IL). The obtained disintegration per minute (dpm) values were corrected for background levels. 78 Miscible Displacement Experiments Miscible displacement experiments were used to determine the hydrodynamic dispersion (Peclet number) and the retardation factor (R) of the column containing nonmodified soil and all the columns containing the HDTMA-modified soil used during this study. A schematic diagram of the miscible displacement system is presented in Figure l. The system consisted of two glass bottles, one containing the solute (the sample) and the second containing elution solvent. Two single piston pumps (Model 302, Gilson Medical Electronics Inc.) with water were used in the constant pressure mode for sample and solvent delivery. The pumps were connected to the column through a four-way solvent selection valve (Model 5020, Reodyne). The inlet of the column was also connected to the four-way solvent selection valve, and the outlet of the column connected to a flowthrough (Holochrome, Gilson) ultra-violet/visible variable-wavelength detector. The column was connected to the system with low dead-volume Teflon fittings and Teflon tubes. The various connections were kept as short as possible to minimize extra- column contributions to dispersion. Changes in concentration as sensed by the detector were registered on a chart recorder (ABB model SE 120). The column for the miscible displacement experiments performed with the native Oshtemo Bt horizon soil was a borosilicate glass preparative column (Beckman, Altex Division) with an intemai diameter of 2.5 cm. The column length was adjusted to 5 cm, using a moveable plunger. Bed supports consisted of woven Teflon (TFE) diffusion mesh and a Teflon (TFE) filter membrane (extra-fine pore-2.5 pm). For the miscible displacement experiments performed with the HDTMA-modified soil, fit-bed columns were used. They were also made from glass borosilicate (Candies, Chromoflex), with an intemai diameter of 2.5 cm., and a fit length of 5 cm. Bed supports consisted of woven FE teflon diffusion mesh and a 2.5-pm Teflon filter membrane together with a 0.45-pm nylon filter membrane. The fraction of air-dried, non-modified or HDTMA- modified soil that passed a 1-mm diameter sieve was used to fill the columns. The 79 528$. :20 new A an: d ESQ? 3083395058? .8 =5me cums—2.0m A Eaufi i meeommo m m < 2:5 9.22.5 e58 111/? ..ow men. ..momHN m mize .5228 .89.... 80 columns were dry-packed using the "rotate, bounce and tap method" (Knox 1976) widely used for packing materials with particle sizes higher than 40 pm. The dry material was packed to obtain a uniform bulk density. For pro—equilibration, the columns were connected, in standing vertical position, to a piston pump (Model 302, Gilson Medical Electronics, Inc.). The non-modified column was equilibrated with a 0.01 N CaC12 solution until steady-state water saturation conditions were established. The modified columns were equilibrated in the same way but with pure water. Columns specification are summarized in Table 2. Table 2. General characteristics of the columns Length (cm) 5 Diameter (cm) 2.5 Input flow rate (ml/min) 3.0 Linear velocity (cm/ hr) 91. 8 To obtain the BTCs in the native Oshtemo Bt horizon soil, a 0.01 N solution of CaC12 filtered through a 0.45 pm Millipore system and degassed with helium was used to prepare solute samples. After determining the sample solution UV deflection (Co) outside of the column at the specific selected wavelength for each solute, the BTC was measured by displacing the sample solution through the column at a linear velocity of approximately 90 cm h'l . The sample was displaced until the detector response reached Co. Then the eluent was switched to 0.01 N CaClz. The column was eluted until the effluent UV response was equal to the original base line. The same basic procedure was used to obtain the BTC through the HDTMA-modified soil column, with the only difference that the solutions were prepared in pure water or different proportions of methanol water mixtures. 81 Benzene BTCs with presmrized system Thesamebasicapparatus (Fig. 1) andthesarnebasicprocedurewereusedtorun miscible displacement experiments through an HDTMA-modified soil column with water and different proportions of methanol and water. However, the solvent-sample delivery system consisted of two glass lined bottles connected through a valve to a He tank which degassed the solutions and maintained the head space slightly pressurized to control evaporation of solvent and indirectly of solute. For the last part of this research, seven columns were gradually equilibrated with increasing proportion of methanol (0 to 40 %) in water and used to obtain benzene BTCs. At the end of each experiment each column was unpacked and the soil air dried and used to determine benzene sorption isotherms, described below. Organic carbon mass balance During the experiments with the HDTMA-modified soil, eluent from each step of the experiment was collected, its volume determined, and the solutions analyzed for l4C-activity in triplicate by. LSC. These data were added together and compared to the specific activity of the soil at the beginning of the experiments and at the end of the experiments. After the experiment, the soils from each of the seven columns were air dried and the soil specific activity and organic carbon content were determined in them. For these analyses, the same procedures described above to obtain the l4C-I-IDT'MA modified soil organic carbon content and specific activity were followed. Sorption Isotherms Batch sorption isotherms from water were measured for solutes (benzene, naphthalene and TCE) in the native Oshtemo Bt soil. For the HDTMA-modified Oshtemo Bt soil, batch sorption isotherms were measured for all solutes in increasing proportions of methanol-water (fc). In addition, benzene sorption isotherms were measured on air-dry HDTMA-treated soil that was removed from the columns at the end 82 of the BTC experiments. Different soil:solution ratios (1:1 to 1:5) were used for the experiment depending upon fc . Larger soilzsolution ratios were used at higher fc in order to improve the precision of the measurement. Duplicate samples were prepared . in 5 ml glass amber tubes with increasing concentrations of14c-labe1ed solute, up to 75 % the water solubility of the solute. The tubes were closed with teflon-lined screw-tops and shaken for 24 h. Following this equilibration period, the solid and liquid phases were separated by centrifugation (1,993 x g) and 1 ml aliquots of the supernatant were analyzed by liquid scintillation counting (LSC). The concentration of solute sorbed in the soil (Q) was determined from the difference between initial and final solution-phase concentrations. Methanol Interfacial Adsorption Equal volumes of octane and water were equilibrated for 48 hours. A series of methanol solutions were prepared in octane-saturated water and equilibrated for four days with an equal volume of water-saturated octane. Using these solutions, the interfacial tension at the octane/water. and air/water interfaces was measured by the drop weight method. The interfacial tension measurement apparatus consisted of a micrometer, syringe, capillary tube, and container. A water solution containing methanol was delivered to the end of one capillary tube with a syringe, with the plunger operated by a micrometer. The tip of the capillary had to be ground smooth so that the end was sharp, regular, fiee from any nicks, and perpendicular to the tube. - A The syringe was previously calibrated to determine the volume of liquid per unit of micrometer scale. As the liquid was delivered from the capillary to the immiscible second phase, a drop formed, and on the break away, the volume of the drop was determined by reading the micrometer. The methanol concentration after equilibrium was determined by gas chromatography (G.C) analysis of 0.5 ml taken from the head space of 10 ml sealed vials containing 1 ml aliquots of the different solutions. The G.C. 83 wasequippedaflameionizationdetectoratatemperatureof 150°C, andaDB—624 column with 3 pm film thickness, 30 m x 0.543 mm at a temperature of 600C. The injection port temperature was 100°C. The densities of water and organic solutions were measured using a picnometer with a 2 ml volume. 4.4 Results and Discussion Estimation of Kw To overcome the experimental limitations expected when dealing with strongly retained volatile organic compounds, the linear relationship between log Km and fc, predicted by the model of Rao et al. ( 1985), discussed in the theory section was used to estimate the sorption coefficients of several solutes from water (Kw). Thus, to obtain Kw, the logarithms of the solutes‘ sorption coefficients obtained from BTCs and batch equilibrium isotherms in mixtures of methanol-water were linearly extrapolated until fo = 0. Five NOCs were used in preliminary tests, benzene, naphthalene, toluene, p- xylene and TCE. Among them, naphthalene had the lowest water solubility (30 ppm) and was therefore most strongly retained in the column. requiring higher proportions of methanol to obtain the BTCs. This allowed use of a wider range of methanol concentrations which provided a more rigorous test of the proposed log-linear relationship between Km and fc. Log Km values for naphthalene as a function of fc, obtained by the two methods as well as the log Kw obtained by the batch method, are presented in Figure 2. The logarithm of Km values estimated from the BTCs decreased linearly as a function of fc. for the range of fc value studied (0.5 to 0.65). A linear equation fitting these data was used to obtain the Kw value of naphthalene, which was ca. 1280. However, the naphthalene Kw value measured directly using the sorption isotherm method (100.9) is approximately one order of magnitude smaller than the value calculated by direct extrapolation assuming a log-linear relationship between Km and fc 2A Naphthalene 5 1 0.9 T 0 . ‘ f I A '0.5 I I I I I ‘ l 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 Methanol fraction I column A batch Figure 2. Log of naphthalene sorption coefficients (K) at different fraction of methanol(fc). 85 over the entire range of methanol concentrations. The Kw values obtained by the two methods correspond physically to the equilibrium distribution constant of the compound between the HDTMA-modified soil and the interstitial water and therefore, their values should be the same. An important observation is that the values of Km obtained by the two methods are very close in value in the range of methanol concentrations 0.5 to 0.65 used to obtain the BTCs, although they do not coincide exactly. In the lower range of methanol concentrations, where Km values were only obtained from sorption isotherm, the log of Km values were not linearly dependent on fc° Negative deviations from linearity were observed below fc = 0.5. Similar results were obtained for benzene and TCE as shown in figures 5 and 6. Deviation from linearity has been observed before with natural soils when acetone has been used as cosolvent (Nkedi-Kizza et al., 1985), with other binary solutions (Morris et al. , 1988), and with reversed phase chromatographic materials (Dorsey and Dill, 1989). To verify the real behavior of the system, BTCs were obtained from water over a complete range of fc values,including pure water (f(3 = 0). BTCs were obtained for two solutes with the relatively high water solubilities (SW), benzene (SW = 1,790 mg/L) and trichloroethylene (TCE) (SW = 1,100 mg/L). The higher water solubilities of benzene and TCE as compared to naphthalene manifests lower K values and enables the determination of BTC in pure water as well as water- methanol mixtures. We were also able to control solute evaporation, a common problem found with volatile compounds, using a new chromatographic commercial pump where the solvent reservoir is pressurized, which minimizes volatilization of the solutes during elution times associated with low fc values (including pure water). In Figures 4 and 5, the BTCs, from water, for benzene and TCE obtained with the HDTMA-modified Oshtemo soil are compared with the BTCs, also from water, obtained for the same compounds in a column filled with the native Oshtemo soil. A shift to the right of the natural Oshtemo soil BTCs is observed in the BTCs of both compounds in HDTMA- modified soil. This shift in the BTCs corresponds to an increase in the R factor for the 86 over the entire range of methanol concentrations. The Kw values obtained by the two methods correspond physically to the equilibrium distribution constant of the compound between the HDTMA-modified soil and the interstitial water and therefore, their values should be the same. 1 An important observation is that the values of Km obtained by the two methods are very close in value in the range of methanol concentrations 0.5 to 0.65 used to obtain the BTCs, although they do not coincide exactly. In the lower range of methanol concentrations, where Km values were only obtained from sorption isotherm, the log of Km values were not linearly dependent on fc. Negative deviations from linearity were observed below fc = 0.5. Similar results were obtained for benzene and TCE as shown in figures 5 and 6. Deviation from linearity has been observed before with natural soils when acetone has been used as cosolvent (Nkedi-Kizza et al., 1985), with other binary solutions (Morris et al. , 1988), and with reversed phase chromatographic materials (Dorsey and Dill, 1989). To verify the real behavior of the system, BTCs were obtained from water over a complete range of fc values,including pure water (fc = 0). BTCs were obtained- for two solutes with the relatively high water solubilities (SW), benzene (SW = 1,790 mg/L) and trichloroethylene (T CE) (Sw = 1,100 mglL). The higher water solubilities of benzene and TCE as compared to naphthalene manifests lower K values and enables the determination of BTC in pure water as well as water- methanol mixtures. We were also able to control solute evaporation, a common problem found with volatile compounds, using a new chromatographic commercial pump where the- solvent reservoir is pressurized, which minimizes volatilization of the solutes during elution times associated with low fc values (including pure water). In Figures 4 and 5, the BTCs, from water, for benzene and TCE obtained with the HDTMA-modified Oshtemo soil are compared with the BTCs, also from water, obtained for the same compounds in a column filled with the native Oshtemo soil. A shift to the right of the natural Oshtemo soil BTCs is observed in the BTCs of both compounds in HDTMA- modified soil. This shift in the BTCs corresponds to an increase in the R factor for the 87 nvocompoundsandmmiscaseindicatesdlatbmzeneandTCEaremoresuongly retained in the HDTMA-modified soil than in the native soil. The R and K values obtained form these BTC experiments are summarized in Table 3. The R factors increased by a factor of about 5 in the HDTMA-modified soil and the K values increased about 15 to 30 times. Some physical and chemical properties of the native and HDTMA-modified Oshtemo soil are given in Table 4. The only apparent changes in the HDTMA-modified soil are a significant increase in the organic carbon content of the soil and a slight decrease (2x) in surface area of the soil. Thus, the increased retention is most likely due to the increase in organic carbon content in the soil produced by the exchange of inorganic cations by HDTMA which manifests higher sorption coefficients and hence larger R values. Table 3. Retardation factor (R) and sorption coefficient (Kw) obtained from BTC. KW obtained from batch sorption isotherm. :arameter 1 Benzene TCE Oshtemo HDTMA-Osh Oshtemo HDTMA- ‘ Osh R 1. 15 6.2 1.36 6.6 Kw(BTC) 0.04 1.3 0.09 1.4 Kw(batch) 0.05 10.005 3.0:t0.09 0.08:1:0.00 2.910.05 5 BTCs and sorption isotherm were experiments used to obtain K’s for benzene and TCE in the HDTMA-treated soil for a range of methanol-water mixtures (fc from 0 to 0.4). Figures 6 and 7 summarize the log of Km values calculated by the two different methods over the fc range. The Kw values obtained fiom the sorption isotherms and the BTCs are different and the relationship between the log of Km and f6 is not a straight 88 1? r AAA A A 0.9“ - A A 0.8“ - A 0.7- ‘ 0.6- ‘ O A - o _ I B 0.5 A 0.4- 0.3- 0.2- _ A 0.1- OJ I I I I I I 0 5_ 10 15 20 25 30 Pore volumes II non-modified soil A HDTMA-modified soil Figure 4. Effect of HDTMA soil modification in benzene breakthrough curve (BTC) in water 89 C/Co O ‘l‘ I b 00' s 10 125 2'0 2'5 8'0 85 Pore volumes I non-modified soil A HDTMA-modified soil Figure 5. Effect of HDTMA soil modification on TCE breakthrough curve (BTC) in water 90 0.5: 0.4- Benzene 0.1 Log K C) i -0.1- _ -0.2- I -0.3- A -0.4- -O.5 l I I JI 0 0.05 0.1 0'15 02 0'25 0.8 0.85 0.4 Methanol fraction (to) II BTCs A sorption isotherms Figure 6. Benzene sorption coefficient (K) obtained from BTCs and sorption isotherms in different fraction of methanol (fc). 91 0.5; 0.4- 0.3‘ 0.2“ 0.1 1 TCE Log K C) -0.1 - -0.2‘ -0.3~ -O.4‘ . Jr -0.5 0.05 01 0'15 012 0.125 03 0.35 0.4 Methanol fraction (fc) I BTC A sorption isotherm Figure 7. TCE sorption coefficient (K) obtained from BTCs and sorption isotherms at different proportions of methanol (fc) 92 line as predicted by the model of Rao’s et al., (1985). It was observed that the organic carbon content in a sample of soil obtained from the column at the end of the miscible displacement experiments was substantially different from the organic carbon content of the freshly modified soil. Difference in the organic ear’oon content of the soil used to obtain the BTC and the sorption isotherms with certainly give different Kw values. Experiments were later designed to study this possibility. Table 4. Properties of the native soil and HDTMA modified Oshtemo soil and of soil columns prepared with these soils. Soil Properties Natural Soil HDTMA-modified Soil Org. Carb. Cont. 0.1 i 0.01 0.8" :t 0.02 (95) Surface Area 5.4 1; 0.1 2.4 :1; 0.08 (111 lg) Column Properties Bulk d 'ty p 1.61 1.60 (slot???)i Poro' 03 - 0.4 0.4 (cm cm ) Organic carbon content prior to elution with water or water-methanol mixtures. The observed deviation from linearity in the relationship between log Km and fc may be related to fundamental differences in the chemical composition of the sorptive phases in native and HDTMA-modified soils. Natural organic matter has a much more polar nature than the HDTMA-derived sorptive phase in the modified soil, which is comprised of the C16 alkyl hydrocarbon matter. Polar functional groups (-COOH, ~OH, -NH2) present in the soil organic matter in either the ionized or nonionized form can interact strongly with water. This interaction decreases the interfacial resistance between the organic matter and aqueous phase and in effect eliminates a well-defined interface. 93 These types of interactions, however, do not likely exist between the non-polar hydrocarbon tailsofHDTMAandthepolarwater, givingrisetoamoredefinedinterface between theHDTMA hydrocarbon tailsandtheintersticial water. Thenatureofthis interface is similar to that between two immiscible liquids (e.g., hexane and water) and that present between water and reverse-phase chromatographic material where hydrocarbons are covalently bonded to the Si-OH functional groups of silica. Amphiphilic molecules such as alcohols have high affinity for these interfaces where they orient according to the polarity of their functional groups (Davis and Rideal, 1961). This creates the possibility of interaction between methanol with the HDTMA-modified soil. Examples of analogous solvent-sorbent interactions have been found in the chromatographic literature (Wahlund, 1979; Scott and Kucera, 1977; Schoenmaher, 1983). Since solvent-sorbent interactions were explicitly ignored in the development of the theory of Rao et al. (1985), interaction of methanol with the organic sorptive phases of the modified soil (sorbent) may explain the deviation of our data from the log-linear relationship predicted by their model. The next section present results of experiments in which we have further explored the possibility of solvent/sorbent interaction. Possible interaction or methanol with the HDTMA on the soil The replacement of inorganic adsorbed cations by HDTMA, which has a long hydrocarbon (C16) chain, separate soil and water by a nonpolar phase. The presence of a sorptive phase comprised of nonpolar hydrocarbon chains that do not interact with water creates a well defined interface between sorbed HDTMA and water. Solute molecules transported to this interface by convective water flow in a column filled with the HDTMA-modified soil will need to cross the interfacial barrier to contact the nonpolar hydrocarbon phase. For HOCs there are two cooperative forces that help them overcome the interfacial barrier: the Van der Waals attractive forces between the solute molecules and the hydrocarbon tail of the HDTMA, and the repulsive hydrophobic force created when a HOC tries to enter in the water and disrupt its hydrogen bond network. 94 In contrast to H003, amphiphilic molecules have a strong tendency to accumulate at the interface, where they can orient themselves with their non-polar part interacting with the hydrocarbon tail and the polar part oriented to the water. This orientation at the interfacedecreases theinterfacialbarrierbetween thewaterandthehydrocarbon phase. Methanol is an amphiphilic molecule and alcohols are well known as surface active compounds (surfactants) (Davies and Rideal, 1961). The interfacial adsorption of methanol could affect the sorption of N OCs by HDTMA-modified soils in methanol- water systems. For this reason, the interfacial adsorption of methanol was studied using the immiscible solvent system octane-water as a model for the HDTMA-water interface in the HDTMA-treated soil. Table 5. Summary of methanol adsorption parameters at the octane-water interface. CMC Am - A G System (%) Azlmolecul Kj / mole e Methanol in 35.8 19.4 5.33 octane-water The results indicate that methanol, as a surface active compound, decreases the interfacial tension between water and octane as a function of concentration (Fig. 8). The decrease in the interfacial tension is nearly linear and follows Henry’s adsorption isotherm. The adsorption parameters of methanol at the octane-water interface obtained from these mesurements are summarized in Table 5. The formation of a single compact layer is observed at a methanol concentration of 35.8 % . The surface area covered (cross sectional) by each molecule at this point is equal to 19.4 A2. The low energy of interaction (5.33 kJ/ mole) indicates that the particles are interacting by London dispersion 95 forces. The concentration of methanol at compact layer is considered to be the critical micelle concentration (CMC). Above the CMC molecules of methanol start to interact with each other in the aqueous phase forming micelles. The observed interfacial adsorption of methanol in an octane-water system could by analogy occur at the HDTMA-water interface in the modified soils. This could affect the sorption of NOCs and in part be responsable for the relationship observed experimentally between the K values for benzene and TCE, and fe One possible mechanism is that the dissolution of methanol into the HDTMA phase, alone or in combination with the adsorption of methanol at the interface could increase the overall polarity of the sorptive phase. This would in turn cause lower sorption coefficients. If we carefully look to all the results obtained for naphthalene, TCE, and benzene during the experiments (Fig. 2, Fig. 5, Fig. 6), two slopes can be discerned rather than the predicted log-linear relationship based on the work by Rao et al. (1985). In the HDTMA-modified soil-water system, the intercept between the two lines of different slopes occurs at fc values between 0.2 and 0.3 (20 to 30 per cent). At fc values above this, the K values are lower than those produced by an extension of the line relating K and fc between 0 and 0.20 methanol. One possible explanation is that at these high methanol. concentrations (> 22 i) the dissolution of methanol into the HDTMA phase could increase the overall polarity of the sorptive phase and hence manifest lower sorption coefficients for NOCs. The observation of an apparent CMC l'or methanol at a concentration of 35.8 96 is also interesting. This suggests a physical change in the methanol-water system that might increase the solubility of the system for NOCs. For example, Kile and Chiou (1989) showed that the solubility enhancement of DDT due to the presence of surfactants is a two-stage process. Below the CMC, the monomeric surfactant slightly increases the solubility of the compound in water; however, above the CMC, a sharp increase in solubility enhancement was observed. With our system the solubility below methanol CMC may be different than above it. Although the CMC of 35.8 % is higher than the 96 A Y‘ (mN/m) I 6 Concentration (M) 1 1O 12 O O ”-4 ‘q 04 Figure 8. Change of interfacial tension 7 in the octane/water interface as a function of methanol concentration 97 intercepts of the experimental lines, it does suggest the possibility of two distinct regions of solubility. At higher fc values relative solvency may increase and account for the change in slope observed experimentally in the plots of log K vs. fc. Organic carbon mass balance The results of the mass balance performed with the 14c determined in solvent collected during the BTC and the organic carbon determined in the columns soil at the end of the BTC are summmarized in Table 6. These results indicate that approximately half of the HDTMA initially present in soil remains adsorbed at the end of each experiment, regardless of methanol concentration. A plot of the amount of HDTMA released from the column at each step of the process (Figure 9) shows that after the equilibration process the amount of carbon eluted from the column is almost zero. These results allow us to conclude that after the equilibration period, a very stable material with an organic carbon content of about 0.4 percent is obtained which corresponds to about half the original HDTMA content. The results also allow us to be sure that the organic carbon content of the soil is the same during the miscible displacement experiments and the corresponding sorption isotherms. 98 Table 6. Summary of organic carbon present in the soil after BTC experiments from an initial organic carbon content in the soil of 0.8 %. 95 left 96 left org.carb.cont. Column from mass from (%) balance org.carb.cont after BTCs 1 63.9 46.2 0.37 :I: 0.001 2 59.9 45.0 0.36 :I: 0.003 3 57.8 46.2 0.37 1; 0.003 4 (") 52.5 0.42 :l: 0.14 5 59.1 45.0 0.36 1; 0.01 6 58.6 45.0 0.36 :1: 0.003 7 53.8 43.8 0.35 i 0.10 (*) loss data Benzene K values from BTCs in pressurized system and sorption isotherms on equilibrated soil The log K values for benzene obtained from BTCs in individual columns with methanol’fractions (fc) ranging from 0 to 0.4 and from sorption isotherms on equilibrated soil removed from the columns after the BTC experiments are presented in Figure 10. As it was demostrated in the previous section, the organic carbon content of the soil was the same during the two determinations of K. However, it is observed in Figure 10 that the values of K obtained by the two methods is different. The gap between the values of K obtained by the two methods (see Fig. 6) has decreased using soils with the same organic but there is still an observable difference. When the equation R = 1 + K p/0 is used to determine K values from the BTC, the conventional approach is to use the bulk density (p), which considers the whole mass of soil as one of the phases with where the compound is distributed. It is assumed re ‘ 16- 8 14- $3 . 1 Water 7,} ,2, 2 5%Meth. < 310%Meth. E ,0- 415%Meth. % 520%Meth. “5 B- 1 630%Meth. o 740%Meth. 5’ 1 *3 6‘ a: 2 o A ‘- 4‘ ‘ g ‘A E, 4 E 6 A ' 7 2.. A AA A A c , r 1.4% o 1000 2000 3000 4000 5000 6000 7000 Desorption Volume (ml) Figure 9. HDTMA desorption during miscible displacement experiments. 100 1 #1 4i 0 - El Sorp.lso. x .5 o BTC -1 .- ‘2 ' I ' I ' I ' I ' 0.0 0.1 0.2 0.3 0.4 0.5 Methanol traetlcn Figure 10. Benzene sorption coefficient (K) estimated from BTCs, obtained in seven columns, and from sorption isotherms obtained in soil from these columns. 101 implicitly that the compound will be uniformly distributed throughout the soil. In the HDTMA-modified soil, the sorption of NOCs will occur in the nonpolar phase created by the HDTMA hydrocarbon tails present on the Oshtemo soil surface. This suggests that the whole soil density will not be the best descriptor of one of the phases where NOCs will be distributed, which means that nature of the phase ratio has changed. However, the K values were obtained from the BTCs were obtained using the traditional phase ratio applied to soil columns. We believe that this change in the nature of the phase ratio is responsible for the difference observed between the two values. However, since we have demostrated the possibility HDTMA/ methanol interaction, calculation of this phase ratio will be a matter of further research. As observed previously, there is a change in slope in the log of Km versus fc at about 22% methanol (0.2), probably explained by the solvent/sorbent interaction. For reverse-phase chromatographic materials, which we believe resemble the HDTMA- modified soil more than a natural soil does, a quadratic relationship between In K and the fraction of organic solvent has been demostrated (Dill, 1987, Dorsey and Dill, 1989; Schoenmakers, et al., 1982). Dill ( 1987) developed a theory, based on lattice statistical thermodynamic theories, to account for retention and selectivity of solutes in reverse- phase liquid chromatography. ln Dill's model the equilibrium constant (K) is given as a simple quadratic function of the permtage of organic solvent. Ying et al., (1989) applied the linearized form of Dill‘s equation to a data base of 346 data sets and found that 80% of the data sets had R2 of 0.9 or higher and almost 50% with R2 of 0.99 or higher. We fitted the benzene BTC data to linear and quadratic equations; the results are presented in Figure 11. A R2 of 0.993 for the quadratic fit vs a R2 of 0.982 for the linear indicates that the HDTMA-modified soil behaves more like a reverse-phase chromatographic material (e.g. C13) than a natural soil. In summary, exchange of native cations in the soil by HDTMA increases retention of NOCs during transport. Sorption coefficients from BTCs run with methanol fractions ranging from 0 to 0.4 were used to demostrate that linear extrapolation of the relationship 102 y . 0345 - 247x - 3.29m R"2 - 0.993 1 . y-o.42-3.79x m2. 0.982 8 .... 9. s: 5 '2 ' I V I '— I ' I ' I 0.0 0.1 0.2 0.3 0.4 0.5 Methanol traetlon Figure 11. Results of fitting benzene breakthrough curve data to a linear and quadratic equation 103 between log Km and fc can not be used to predict sorption coefficients of NOCs from water in the HDTMA-modified soil because the relationship deviates from linearity below 30 % methanol. Interaction between methanol and HDTMA is probably responsible for this behavior. It was also found that benzene sorption coefficients determined directly from water using BTCs and sorption isotherms are different. We believe that these values are different because I(W is calculated from the BTC using the phase ratio (p/O) applied to natural soils. In the modified soil, the nature of the phase ratio appears to be different, although it is difficult if not impossible to calculate with the present level of knowledge. The relationship between log K vs fc fits a quadratic better than a linear equation, reinforcing our belief that the modified soil resembles more a reverse-phase chromatographic material than a natural soil. 104 References Amidon, G.L., S.H. Yalkosky, and S. Leung. 1974. Solubility of nonelectrolytes in 1133?; solvents ll: solubility of aliphatic alcohols in water. J. Pharm. Sci. 63, Axilrod, R.M. and E. Teller. 1943. Interaction of the van der Waals type between three atoms. J. Chem. Phys. 11, 299. Boyd S.A., J-F. Lee, and M.M. Mortland. 1988. Attenuating organic compound mobility by soil modification. Nature. 333, 345. Davis, J.T:7and E.K. Rideal. Interfacial phenomena. Academic Press, New York, 1961 p 4 4. Dill, K.A. 1987 . The mechanism of solute retention in reversed-phase chromatoraphy. J. Phys. Chem. 1980-1988 Dorsey J .G. and K.A. Dill. 1989. The molecular mechanism of retention in reversed- phase liquid chromatography. Chem. Rev. 89, 331. Halicioglu, T. and O. Sinanoglu. 1969. Solvent effects on cis-trans azobenzene isomerization: A detailed application of a theo of solvent effects on molecular associations. Ann. New York Academic Sci. 1 8, 308. Hildebrand, J .H., J .M. Prausnits, and R.L. Scott. Regular and related solutions. Van Nostrand, New York, 1970. Horvath, C. , W. Melander and I. Molnar. 1976. Solvophobic interactions in liquid chromatography with nonpolar stationary phases. J. Chrom. 125, 129. Karickhogf, SW. 1984. Pollutant sorption in aquatic systems. J. Hydraulic Eng. 110, 7 7. Kestrler, N.R. and O. Sinanoglu. 1962. Effective intermolecular pair potentials in nonpolar media. J. Chem. Phys. 38, 1730. Kile, D.E. and CT. Chiou. 1989. Water solubility enhancements of DDT and trichlorobenzene by some surfactants above and below the critical micelle concentration. Environ. Sci. Technol. 23, 832. Knox, J .H. 1976. Practical high performance liquid chromatography. Ed. Simpson, Heyden & Son Ltd. , London. Lee, 1.1=., J.R. Crum, and S.A. Boyd. 1989. Environ. Sci. Technol. 23, 1365. McCormick, R.M. and B.L. Karger. 1980. Distribution phenomena of mobile-phase components and determination of dead volume in reversed-phase liquid chromatography. Anal. Chem. 52, 2249. Morris, K.R.; R. Abramowitz, R. Final, P. Davis, and S.H. Yalkowsky. 1988. Solubility of aromatic pollutants in mixed solvents. Chemosphere. 17, 285. 105 Nkedi-Kim, P., P.S.C. Rao, and A.G. Homsby. 1985. Influence of organic cosohlll‘reiltslgn évigigption of hydrophobic organic chemicals by soils. Environ. Sci. Tee 0 . , . Nkedi-Kim, P., P.S.C. Rao, and A.G. Homsby. 1987. Influence of organic cosolvents on leachin87of hydrophobic organic chemicals through soils. Environ. Sci. Technol. 21, 11 . - Nogare, SD. and J. Chiu. 1962. A study of the performance of packed gas chromatography columns. Anal. Chem. 34, 890. Nye, J.V. Biocompatibility of quaternary ammonium cations used to enhance the imlgobilization of nonionic organic contaminants in soils. M.Sc., Thesis. 1993. M. .U. Rao, P.S.C., A.G. Hornsby, D.P. Kilcrease, and P. Nkedi-Kim. 1985. Sorption and transport of hydrophobic organic chemicals in aqueous and mixed solvent sys6tems: Model development and preliminary evaluation. J. Environ. Qual. 14, 37 . Schoenmakers, P.J., H.A.H. Billiet and L. de Galan. 1983. Description of solute detention over the full range of mobile phase com sitions in reversed-phase liquid chromatography. J. Chromatography, 282, l . Scott, R.P.W. and P. Kucera. 1977. Examination of five commercially available chromatographic reversed phases (including the nature of the solute-solvent- stationary phase interactions associated with them. J. Chromatography, 142, 213. Slaats, E.H., W. Markovski, J. Fekete and H. Poppe. 1981. Distribution equilibria of solvent components in reversed-phase liquid chromatographic columns and relationship with the mobile phase volume. J. Chromatography, 207, 299. Wahlund, K-G. 1975 . Reversed-phase ion-pair partition chromatography of carboxilates and sulphonates. J. Chromatogr. 115, 411-422. Wekhove-Goewie, C.E., U.A.Th. Brikman and R.W. Frei. 1981. Trace enrichment of polar compounds on chemically bonded and carbonaceous sorbents and application to chlorophenols. Anal. Chem. 53, 2072. Wood, A.L., D.C. Bouchard, M.L. Brusseau, and P.S.C. Rao. 1990. Cosolvent effects on sorption and mobility of organic contaminants in soils. Chemosphere. 21 , 575. Yalkowsky, S.H., G.L. Amidon, G. Zografi, and G.L. Flynn. 1975. Solubility of nonelectrolytes in polar solvents lll: alkyl p-amonobenzcates in polar and mixed solvents. J. Pharm. Sci. 64, 48. Yaynes, W.F. and S.A. Boyd. 1991. Clay mineral type and organic sorption by hexadecyltrimathylammonium-exchanged clays. Soil Sci. Soc. Am. J. 55, 1991. Ying, P.T.,J.G. Dorsey, and K.A. Dill. 1989. Retention mechanisms of reversed-phase liquid chromatography: Determination of solute-solvent interaction free energies. Anal. Chem. 61, 2540-2546. MICHIGAN $7an UNIV. LIBRnRIEs illWIN“liIUIIHWIiIWII”lllllllllllllllllllmlll 31293010330409