1' < ,L- tm'" ‘ . ‘ Agni?"- .‘~‘ “9?" 17 r w ‘J ‘ .“i-i‘b'i‘xl‘ "J. . “alum” ‘ in L! ,V " pm: - -.....A y“ .1“ '0 ’1' -- .-z . . I I 7 Big ("I In: ’2 H' 5;? +17.” . $3,? $3 ‘3 ‘ “.mg’." g: @233”! r A W; mfg! P Jr: .3 2' ea? ,3 ‘ 1 f 'r ‘1‘: '4 "gr 11? g L; ”9“" LIBRARY Michigan State University iw This is to certify that the dissertation entitled INTEGRATING MOLECULAR SIMULATIONS WITH EXPERIMENTS TO STUDY ORGANIC POLLUTANT INTERACTIONS WITH CLAY MINERALS presented by VANEET AGGARWAL has been accepted towards fulfillment of the requirements for the Ph.D. degree in CROP AND SOIL SCIENCE ‘WXW Major Wofessor’s Signature 2? (Italy)— 7/00 3 U Date MSU is an Affirmative Action/Equal Opportunity Institution PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE J7 DATE DUE DATE DUE W 3.35 21% 2/05 c:/ClRC/DmDm.indd-p.1s -i_____ INTEGRATING MOLECULAR SIMULATIONS WITH EXPERIMENTS TO STUDY ORGANIC POLLUTANT INTERACTIONS WITH CLAY MINERALS By VANEET AGGARWAL A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Crop & Soil Science 2005 ABSTRACT INTEGRATING MOLECULAR SIMULATIONS WITH EXPERIMENTS TO STUDY ORGANIC POLLUTANT INTERACTIONS WITH CLAY MINERALS By VANEET AGGARWAL Triazine herbicides such as atrazine and simazine have been applied for more than 40 years in the United States. Atrazine is one of the most common herbicide residues found in groundwaters of Midwest. Sorption to soil components can control the bioavailability, persistence, and transport of xenobiotics. We investigated the sorption of three triazine herbicides (atrazine, simazine, and metribuzin) by saponite and beidellite clay minerals saturated with K+, Cs+, Na+, and Ca“. Saponite clay sorbed a larger fraction of each pesticide from aqueous solution than did beidellite clay. The lower cation exchange capacity in saponite (vs. beidellite) presumably results in a less crowded interlayer with more siloxane surface available for adsorption. Generally, Cs-saturated clays sorbed more triazines than clays saturated by K+, Na+ or Ca2+. We attribute this to the smaller hydrated radius of Cs+, which again increases the siloxane surface available for sorption. Furthermore, the relatively weak hydration of Cs+ reduces the swelling of clay interlayers, thus making sorption domains less hydrated and more receptive to hydrophobic molecules. Cs-saponite manifested a sorption of more than 1% atrazine by weight above equilibrium concentrations of 6 mg/L. This is the largest sorption of neutral atrazine from water yet reported for an inorganic sorbent. Molecular dynamics simulations indicate that atrazine interacts with both clay basal planes and with multiple cations in the clay interlayer forming both inner- and outer-sphere complexes. Similar effects of exchangeable cation on trichloroethene sorption by saponite clay were also observed, and our general concept of the factors controlling sorption of nonpolar organics by clay minerals is becoming more clear. Molecular simillations were also used to estimate thermodynamics of polar organic sorption to montmorillonites. Our method shows promise in that simulated adsorption enthalpies for nitroaromatic compounds have the correct Sign and are significantly different from that of p-xylene. Further work is still needed so that enthalpy predictions can be more precise. DEDICATION To my daughter, Muskaan, wife Deepa, and parents, for everything iv ACKNOWLEDGEMENTS I would especially like to thank my advisor, Dr. Brian Teppen, for his guidance, support, and encouragement throughout my program. I have enjoyed our interactions immensely. I thank Dr. Teppen for providing me with countless opportunities to grow both personally and professionally. This dissertation could not have been completed without his help and support. I thank my guidance committee, Dr. Stephen Boyd, Dr. Donald Penner, and Dr Thomas Voice for their insight and helpful comments. I would like to offer my gratitude to numerous people, especially, Dr. Hui Li for helping me with the laboratory work, Michael Roberts for being there whenever I would need him, Simone, Jacqui, Jim, Chunhui, Elica, Dr. Yao-Ying, and several others, who have encouraged and supported me. I would also like to acknowledge the Dissertation Completion Fellowships provided by the Environmental Science and Policy Program and by College of Agriculture and Natural Resources, Kirk and Marjorie Lawton for the graduate student support, chair of Crop & Soil Sciences department for the Retention Fellowship, and Punjab Agricultural University for the study leave to pursue my dream of getting PhD. in the US. Finally, I have been truly blessed to have the love and support of my wonderful and encouraging family, my wife Deepa, our precious daughter Muskaan, my parents, my sister Dimple and her husband Ajay. Their encouragement, understanding, prayers, and being here on my defense and graduation are much appreciated. This accomplishment is as much theirs as it is mine, and to them I am forever thankful and grateful. TABLE OF CONTENTS LIST OF TABLES viii CHAPTER I: INTEGRATING MOLECULAR SIMULATIONS WITH EXPERIMENTS TO STUDY ORGANIC POLLUTANT INTERACTIONS WITH CLAY MINERALS .................................................................... 1 INTRODUCTION ............................................................................. 2 SORPTION OF ATRAZINE TO SOIL COLLOIDS ......................................................... 6 PROPOSAL ............................................................................................................ 13 LITERATURE CITED ............................................................................................. 22 CHAPTER H: TRIAZINE ADSORPTION BY SAPONITE AND BEIDELLITE CLAY MINERALS ............................................................................ 29 ABSTRACT ............................................................................................................ 30 INTRODUCTION .................................................................................................... 31 MATERIALS AND METHODS ................................................................................ 34 RESULTS AND DISCUSSION ................................................................................. 4O LITERATURE CITED ............................................................................................. 55 CHAPTER III: ATRAZINE INTERACTIONS WITH SMECTITE CLAYS: MOLECULAR SIMULATION APPROACH .......................................................... 59 INTRODUCTION .................................................................................................... 60 MATERIALS AND METHODS ................................................................................ 65 RESULTS AND DISCUSSION ................................................................................. 69 LITERATURE CITED ............................................................................................ 86 CHAPTER IV: ENHANCED SORPTION OF TRICHLOROETHENE BY SMECTITE CLAY EXCHANGED WITH CS+ ...................................................... 91 ABSTRACT ........................................................................................................... 92 INTRODUCTION ................................................................................................... 93 MATERIALS AND METHODS ............................................................................... 98 RESULTS AND DISCUSSION ............................................................................... 103 LITERATURE CITED ........................................................................................... 1 13 vi CHAPTER V: MOLECULAR SIMULATIONS TO ESTIMATE THERMODYNAMICS OF POLAR ORGANIC SORPTION TO MONTMORILLONITE ........................................................................................... 1 18 INTRODUCTION .................................................................................................. 1 19 MATERIALS AND METHODS .............................................................................. 124 RESULTS AND DISCUSSION ................................................................................ 134 LITERATURE CITED ........................................................................................... 149 vii LIST OF TABLES CHAPTER II Table 1. Selected Physicochemical Properties of Pesticides Used in this Sorption Study ..................................................................................... 35 CHAPTER V Table 1. Estimated and measured enthalpies of reaction (kJ/mol) as defined by the equation designated as cycle [2] and Figure 2. The same estimates for AH; and AH5 were used for each compound ......................................................... 135 Table 2. The F reundlich adsorption coefficients for substituted nitrobenzenes and the adsorption energies of the corresponding substituents .......................... . 137 Table 3. Application of algorithm 2 to estimate experimental adsorption enthalpies ....................................................................................................... 1 47 viii LIST OF FIGURES CHAPTER I Figure 1. Schematic for decomposing the energy changes for transfer of a compound from aqueous solution to the clay layer (after G033 and Schwarzenbach, 2001) ...................................................................................... 19 CHAPTER II Figure 1. Charges on atrazine molecule used in simulation study ............................. 39 Figure 2. Sorption isotherms representing atrazine (a,b), metribuzin (c,d), and simazine (e,t) uptake from water by reference saponite (SapCa-2) and beidellite (SBId-l) clays saturated with Cs+, K+, Na+ and Ca2+ ...................... 42 Figure 3. Comparison of our sorption results with all earlier studies of atrazine sorption to saponites, as well as two representative montmorillonites ............ 44 Figure 4. X-ray diffraction patterns of air-dried Cs-SapCa-2 clay films as a function of atrazine sorption ............................................................................. 50 Figure 5. Snapshot of a molecular dynamic Simulation of atrazine in the interlayer region of homoionic K-smectite (K-SapCa—Z). Clay layers were parallel to the page and have been removed in order to expose the interlayer complexes. The atom coloring scheme is K = green, O = red, H = white, C =gray, N =b1ue, and C1 = green (associated with the atrazine molecule) ..... 52 CHAPTER III Figure 1. A typical starting configuration before beginning the molecular dynamics simulation. Here, atrazine has been placed into the interlayer of Cs-SWy-Z with 55 water molecules. The equilibrium dam-Spacing will eventually be 12.3 A ........................................................................................ 67 Figure 2. Partial atomic charges on the atoms of the atrazine molecule, as used in all simulation studies. The total charge on the molecule sums to zero ........ 68 Figure 3. Snapshot showing the orientation of atrazine within the interlayer of a K-saturated smectite like SWy-2 containing (A) one, (B) two, and (C) three layers of hydration water (red = oxygen, white = hydrogen, blue = nitrogen, gray = carbon, green = interlayer K+ cations). Images on the left depict a side-view (x,z plane) of the atrazine-smectite interlayer complex while the images on the right depict the same from a top—view (x,y plane), having removed the silicate layers. The simulated door-spacings for A, B, and C are 12.4, 15.5, and 17.3 A, respectively ............................. 71-72 Figure 4. Snapshots of molecular dynamic simulations of atrazine in the interlayer regions of four different homoionic smectites, (A) K-SWy-Z, (B) Cs-SWy-2, (C) K-SAz-l, and (D) K-SapCa-2. Clay layers were parallel to the page and have been removed in order to expose the interlayer complexes. The atom coloring scheme is K = green, 0 = red, H = white, C =gray, N =b1ue, and C1 = green (associated with the atrazine molecule) ......................................... 73 Figure 5. Trajectory of the Cs -- Cl inner sphere nonbonded complex in the Simulated atrazine-Cs-SWy-Z system represented by Figure 4B .................... 74 Figure 6. Representative snapshot of the relationships of atrazine and cations to charged sites in model SWy-2 with a monolayer of water for the K- (left) and Cs-saturated (right) cases. Blue octahedra are the locations of Mg-for- Al octahedral substitution sites and pink tetrahedral are the locations of Al- for-Si tetrahedral substitution sites in the montmorillonite layer. The upper clay layer and the rest of the lower clay layer have been removed for clarity ................................................................................................................ 78 Figure 7. Snapshots of molecular dynamic Simulations of atrazine in the interlayer regions of four different homoionic smectites, (A) K-SAz-l, (B) Ca-SAz-l, (C) K-SWy-2, and (D) Cs-SWy-Z. Clay layers were parallel to the page and have been removed in order to expose the interlayer complexes. The atom coloring scheme is K = green, 0 = red, H = white, C =gray, N =b1ue, and C1 = green (associated with the atrazine molecule) ................................... 79 Figure 8. Trajectory of the K -- N inner sphere nonbonded complex in the simulated atrazine-K-SWy-Z system represented by Figure 3B and 7C .......... 80 Figure 9. Representative snapshots of the relationships of atrazine and cations to charged sites in model SWy-Z with a bilayer of water for the K- (left) and Cs-saturated (right) cases. Blue octahedra are the locations of Mg-for- Al octahedral substitution sites and pink tetrahedral are the locations of Al- for-Si tetrahedral substitution sites in the montmorillonite layer. The lower clay layer and the rest of the upper clay layer have been removed for clarity ................................................................................................................ 81 Figure 10. Snapshots of molecular dynamic simulations of atrazine in the interlayer regions of two different homoionic smectites, (A) C a-SAz-l , and (B) K-SWy-2. Clay layers were parallel to the page and have been removed in order to expose the interlayer complexes. The atom coloring scheme is K = green, 0 = red, H = white, C =gray, N =b1ue, and C1 = green (associated with the atrazine molecule) ........................................................... 83 CHAPTER IV Figure l. Sorption isotherms representing trichloroethene uptake from water by reference saponite (SapCa-2) clay saturated with (a) calcium, and potassium, and (b) cesium ............................................................................................... 105 Figure 2. X-ray diffraction pattern of reference saponite (SapCa-2) clay in oriented clay films as a function of different cation saturation at 100% humidity ...... 109 CHAPTER V Figure 1. Charges on para-dinitrobenzene molecule used in simulation study ......... 126 Figure 2. Schematic for decomposing the energy changes for transfer of a compound from aqueous solution to the clay layer (afier G053 and Schwarzenbach, 2001) ................................................................................... 128 Figure 3. Validity of algorithm with experimental data ............................................ 144 Figure 4. Sensitivity Analysis .................................................................................... 146 “Images in this thesis/dissertation are presented in color.” xi CHAPTER I INTEGRATING MOLECULAR SIMULATIONS WITH EXPERIMENTS TO STUDY ORGANIC POLLUTANT INTERACTIONS WITH CLAY MINERALS Introduction Atrazine is one of the most commonly used agricultural herbicides in the world and has been applied for more than 30 years by com and sorghum growers in the United States (Ribaudo and Bouzaher, 1994). In 1994, over 47 million pounds of atrazine were applied nation-wide covering about 77 % of the area cropped under com-soybean rotation (Pike and Knake, 1996). About 2.4 million pounds of atrazine were applied on 80 0/0 of the cropped area in Michigan, making it the most widely used herbicide in Michigan (Pike and Knake, 1996). Atrazine is also one of the most common herbicide residues found in groundwaters of Midwest (Blanchard and Donald, 1997; Burkart et al., 1999; Kolpin et al., 1998). Atrazine and two of its metabolites were detected in 26 % of 303 sampled Midwest wells in the United States (Kolpin et al., 1996) and is detected 10 to 20 times more frequently than the next most commonly detected pesticide (Belluck et al., 1991). Based on 41 land-use studies across the United Sates during 1993-1995, (Kolpin et al., 1998) indicated that of the 46 pesticide compounds examined, 39 were detected in shallow groundwater. Atrazine was the most frequently detected (38 0/0) followed by its decomposition product, deethylatrazine (34 %). The presence of atrazine in groundwater is a concern because of potential carcinogenic effects ofs-triazines, which include atrazine, simazine, and propazine (Biradar and Rayburn, 1995). Other toxic effects of atrazine include skin irritation, respiratory effects, central nervous system disorders, and blood cell disorders (US. Congress, Office of Technology Assessment, 1984). Adsorption and desorption processes on soil components are important phenomena that control the environmental fate and behavior of pesticides including bioavailabilty, persistence and transport of pesticides in soils and subsoils. Clay minerals and soil organic matter (SOM) are the two chemically most active components of the soils. Generally, cultivated soils contain a larger fraction of clay than organic matter, yet most of the research conducted in the last 20 years has considered SOM as the primary sorptive phase of organic contaminants interacting with soils (Chiou et al., 1979; Chiou et al., 1983; Chiou et al., 2000; Huang et al., 1997; Kleineidam et al., 1999; Weber and Huang, 1996; Xia and Ball, 2000; Xing and Pignatello, 1997). Chiou et a1. (1979) in a study with seven neutral organic molecules did not observe any indication of adsorption isotherm curvature even at concentrations approaching saturation. Therefore, they concluded that adsorption of neutral organic molecules is controlled by soil organic matter and occurs essentially by partition processes. Chiou et a1. (1983) even suggested that sorption of non-ionic organic compounds on the soil mineral fraction is relatively unimportant in moist soils. Xing and Pignatello (1997) in a study with chlorinated benzenes, observed that the value of Freundlich coefficient, N, was unaffected by complete removal of mineral phase, again indicating that soil organic matter controls the sorption of such compounds in soils. It has also been found that in addition to the organic carbon content, the chemical and structural properties of soil organic matter also significantly affect the sorption of Slightly polar and non-polar hydrophobic organic contaminants (Grathwohl, 1990; McGinley et al., 1993; Weber et al., 1992; Young and Weber, 1995). Geologically mature organic matter had much higher sorption capacities than relatively young humic organic matter in soils (Grathwohl, 1990; Young and Weber. 1995). Kleineidam et a1. (1999) in a study with low organic matter soils observed that the reduction of organic carbon significantly decreased sorption. They therefore concluded that organic matter dominates sorption of hydrophobic organic compounds, even in samples with very low organic carbon contents. Clay minerals are layered aluminosilicates, most of which have structural negative charges due to cation substitution and surface areas exceeding 105 in2 kg71 (Borchardt, 1989; Brown et al., 1978). The large surface area per unit weight associated with clay minerals is important for sorption of pesticides (Karickhoff, 1984). Among commonly found clay minerals, expandable 2:1 layer silicate clays are especially important due to their high surface areas, high cation exchange capacities (C EC 3) and surface reactivities (McBride, 1994). Smectites are important members of the expandable 2:1 layer silicate clays and contribute much of the inorganic surface areas of many soils. Therefore, smectites have a high potential for influencing the fate of pesticides in the soil environment. The sorption capacity of smectites for atrazine varies widely depending on the surface charge density of the smectite (Laird et al., 1992), the nature of the adsorbed cation (Sheng et al., 2001), and the pH of the soil (Vaz et al., 1997) Hayes and Mingelgrin (1991) reported that in low organic matter soils and sediments, clays might control the sorption/desorption interactions. Khan et al. (1979) observed that inorganic constituents could have a dominant role in pesticide retention in low organic matter soils. Means et al. (1982) reported that pesticide sorption is proportional to the percent organic carbon and where percent organic carbon is low, the amount of clay becomes important in determining pesticide sorption. Karickhoff (1984) concluded that at clay to organic matter ratios greater than 30, the soil mineral phase becomes important for the organic contaminant sorption. Rebhun et al. (1992) reported that sorption of organic contaminants by the organic fraction was 8 to 20 times higher than by pure clay minerals. However, in soils with the commonly encountered range of organic matter (0.5 — 5 %) and high clay content, the contribution of the mineral fraction exceeded that of the organic component. In a study on the adsorption of some nitroaromatics by clay minerals, Haderlein and Schwarzenbach (1993) and Haderlein et a]. (1996) found that clays with high surface areas (e. g., smectites) and exchangeable cations with low hydration energies (e.g., Kr and CS+) promoted adsorption. Boyd et al. (2001) demonstrated that the strong sorption of nitroaromatiCs occurs because the organic compound forms interlayer complexes with exchangeable cations of the clay. Recently, Sheng et a1. (2001) reported that on a unit mass basis, some smectite clays were more effective sorbents than SOM for four out of the seven pesticides studied. Hance (1965) suggested that under field conditions in the absence of excess water, the soil mineral fraction may play an important role in adsorption. Similar results were observed for adsorption of parathion (Yaron and Saltzman, 1972), and parathion and lindane (C hiou et al., 1985) on soils. In a dry soil—hexane—parathion system the Slightly polar parathion molecules efficiently compete with the apolar hexane molecules for adsorption Sites (Yaron and Saltzman, 1972). Generally as the soil water content increases, parathion adsorption decreases because of the decreasing contribution of the soil mineral fraction. When soils are fully hydrated, adsorption of neutral organic solutes by soil minerals often becomes relatively insignificant compared to the uptake by partitioning into soil organic matter, presumably because water is preferentially adsorbed by minerals (Chiou et al.. 1985). Sorption of atrazine to soil colloids a. Organic matter and clays Talbert and F letchall (1965) showed that atrazine adsorption to whole soils was strongly correlated with both clay content and organic matter content. However, Roy and Krapac (1994) did not find any correlation between atrazine sorption and clay content. Terce and Calvet (1977) reported little atrazine sorption by Ca-smectites. Borggaard and Streibi g (1988) reported that smectites and iron oxides adsorbed little or no atrazine in the pH range of 4 — 8 and suggested that organic matter is possibly the only atrazine- adsorbing constituent in many natural soils. Brouwer et al. (1990) observed a logarithmic increase of atrazine distribution coefficient with organic matter content. Wang et a1. (1990) repOrted the preferential binding of atrazine to the higher molecular weight fraction of fulvic acids and proposed weak mechanisms of interaction such as hydrogen bonding and hydrophobic bonding to interpret their findings. On the basis of NMR spectroscopic investigation, Welhouse and Bleam (1992) concluded that due to the separation of charge within the atrazine molecule, atrazine is able to both donate and accept hydrogen bonds. On the basis Of intermediate values of atrazine’s donating (am). and accepting (13m) parameters, they suggested that particularly strong complexes could be expected between atrazine and acid functional groups of humic substances (Welhouse and Bleam, 1993a) and concluded that hydrogen bonding is an important mechanism for atrazine adsorption to soil organic matter (Welhouse and Bleam, 1993b). Grundl and Small (1993) evaluated the role of mineral-phase sorption of atrazine and alachlor by different sediments and concluded that they were sorbed by both natural organic carbon (OC) and clays. The critical clay/OC ratios at which mineral-phase sorption accounted for 50 % of the overall sorption was 62 and 84 for atrazine and alachlor, respectively. In a study with a Webster soil, Laird et al. (1994) found that organic and inorganic components were 11 and 89 % of the mass and contributed 68 and 32 "/0 respectively, of the affinity for atrazine. Laird et al. (1992) worked in the low-concentration range up to 0.04 mg atrazine/ g clay and reported that atrazine adsorption by 14 different Ca-smectites ranged from 0 to 100 %. Recently, Sheng et al. (2001) showed that a K-Saturated smectite and a muck soil sorbed equal amounts of atrazine from aqueous solution. b. Exchangeable cations The type of exchangeable cation on the clay also affects pesticide sorption. Sheng et al. (2001) reported that homoionic K-montmorillonite adsorbed up to 240 mg atrazine/kg clay and is seven times more effective sorbent for atrazine than homoionic Ca-montmorillonite. They reported that the lower effectiveness of Ca-montmorillonite might be due in part to the higher hydration energy and hence larger hydrated radius of Ca2+ than that of K+. Thus the waters of hydration associated with Ca2+ obscure a greater portion of the clay surface than those associated with K+, which in turn will shrink the size of the adsorptive domains between exchangeable cations. They reported that 12.5 A basal spacing associated with many K-smectites, as opposed to >15 A spacing for C a- smectites, may be optimal for adsorption of semiplanar aromatic pesticides like atrazine. Sawhney and Singh (1997) also reported higher atrazine adsorption by Al-smectite (3.82 mg atrazine/ g clay) than by Ca-smectite (1.90 mg atrazine/ g clay). They attributed higher atrazine sorption by Al-Smeetite to the more polarized water molecules associated with Al3+ than with Ca2+; those water molecules are postulated to electrostatically attract N- groups of atrazine. Macnamara and Toth (1970) observed that K-clay systems Showed a greater adsorption capacity for linuron and malathion than Ca-, Mg-, and H/Al- clay systems. Using on-line microfiltration and HPLC technique, Gilchrist et al. (1993) observed that most of the added atrazine was rapidly and reversibly sorbed on the montmorillonite and that Na+ saturated montmorillonite sorbed more atrazine than Ca2+ saturated montmorillonite. The Ca2+ saturated montmorillonite had a higher surface charge density than Na+. c. Surface charge density (SC D) Laird et al. (1992) worked in the low-concentration range up to 0.04 mg atrazine/ g clay and reported that atrazine adsorption by 14 different Ca-smectites ranged from 0 to 100 %. They reported that Panther Creek beidellite adsorbed virtually all of the added atrazine while Otay, CA montmorillonite did not adsorb any atrazine. The amount of atrazine adsorbed by the remaining 12 smectites varied between these two extremes. Panther Creek beidellite had the lowest CBC (88 cmol kg‘l) while Otay, CA montmorillonite had the highest CEC (134 cmol kg'l), therefore they concluded that atrazine adsorption is inversely related with the CEC of the smectite clays. They found that 83 “/0 of the variation in the values of the Freundlich constant (log K) was related to the surface charge density (SCD) and surface area (SA) of the clays. Panther Creek beidellite had lower values of SCD (1.42 umol m'z) and SA (619 m2 g") while Otay, CA montmorillonite had higher SCD (1.85 umol m'z) and SA (724 m2 g") values. In a study with six different species of K-smectites, Sheng et al. (2001), reported that the amount of atrazine adsorbed by K-saponite was 0.25 mg/ g clay and by K-montmorillonite was 0.10 mg/ g clay. On the basis of adsorption coefficient values (Kr), they concluded that K- saponite (CEC 94.9 cmol/kg) is about 15 times more effective in the removal of atrazine from the solution than K-montmorillonite (CEC 130 cmol/kg). These observations are in agreement with the results of Laird et a1. (1992) who concluded that atrazine sorption by clays is inversely dependent on clay properties such as SCD and CEC. Based on this evidence, they concluded that under neutral pH conditions, atrazine is dominantly sorbed on smectites as molecular species. Hamaker and Thompson (1972) suggested hydrogen bonding as a mechanism for the sorption of s-triazine herbicides to clay surfaces. Laird (1996) reported that under neutral conditions, molecular atrazine is initially sorbed on smectites by a combination of water-bridging and hydrophobic interactions. He proposed that the lone pair of electrons on the ring N atoms of atrazine molecules interact with water molecules solvating metal cations adsorbed on smectite surfaces. At the same time, the alkyl-side chains of atrazine molecules out-compete water molecules for retention on hydrophobic microsites on smectite surfaces. Hydrophobic bonding is also consistent with the observed decrease in sorption with increasing surface charge density of smectites because the amount of hydrophobic surface on smectites decreases with increasing surface charge density. One reason for this decrease in sorption with increase in SCD is that greater SCD causes the water molecules at the surface of the clays to be more strongly bound and have a more ordered structure (Gilchrist et al., 1993). Since atrazine is only sparingly water soluble (water solubility is only 30 mg/L), it is less likely to approach and adsorb to surfaces having rigid, ordered water layers than to surfaces without such layers. Barriuso et a1. (1994) reported that in spite of the lower adsorption capacity of high SCD clays, atrazine was held by stronger binding mechanisms on these clays. d. pH The extent of pesticide adsorption by clays depends strongly on the molecular structure including the basicity of the pesticides and thus on the pH of the solution, with maximum adsorption typically occurring at a pH near the pKa values of the pesticides (Weber, 1970). Atrazine has pKa of 1.68, so maximum adsorption of atrazine is expected in very acidic soils. Sorption of atrazine on acidic humic acid (pH 2.5) was ten times greater (Kd 627 cm3 g'l) than on neutral (pH 7.0) humic acid (Kd 62.2 cm3 g") (McGlamery and Slife, 1966). This pH effect was attributed to increased ionic bonding caused by protonation of the amino groups on the atrazine molecule at low pH (McGlamery and Slife, 1966). Weber (1970) concluded that the adsorption of s-triazines was due to complexation of the triazine molecules with functional groups on the organic colloids and/or adsorption of s-triazine cations by ion-exchange forces. Gilmour and Coleman (1971) showed that cation exchange causes a large increase in atrazine sorption to humic acid at low pH. Charge transfer complexes have also been postulated as a possible mechanism involved in the adsorption of s-triazine herbicide onto SOM (Hamaker and Thompson, 1972). Borggaard and Streibig (1988) reported that smectites and iron oxides adsorbed no or insignificant atrazine in the pH range of 4 — 8. Madsen et a1. (2000) reported that sorption of atrazine depends primarily on pH, the sorption being high at low pH (<6.7) compared with sorption at higher pH (>7.4). At lower pH, more hydrogen ions associate with the atrazine molecules giving more cationic characteristics, which leads to increased adsorption (Frissel and Bolt, 1962). Baskaran et al. (1996) reported that as the pH increases, the proportion of atrazine cations in the solution decreases and hence the decrease in sorption with increase in pH. Clay et al. (1988) also reported that with the application of ammonium sulfate and elemental S as fertilizers to the soils, the adsorption of atrazine increased as the fertilizers decreased soil pH from 6.0 to 4.5. Atrazine is an extremely weak base (pKa of 1.68), hence little protonation of adsorptive phases can be expected at pH values 2 or more units above the pKa of the adsorbate (Hayes and Mingelgrin, 1991). Therefore, atrazine should be sorbed primarily as a neutral molecule in cultivated soils, which have a pH range of 4 — 8. Laird et al. (1992) also reported that atrazine is dominantly sorbed on smectites as a neutral species for the pH range 4.75 to 6.45. e. Temperature Harris and Warren (1964) reported that adsorption of atrazine, monuron, and simazine by bentonite was greater at 0°C than at 50°C. Adsorption by muck was similar at the two temperatures for all compounds. They concluded that bentonite had a higher bonding enthalpy than muck. Talbert and Fletchall (1965) also reported that the amount of atrazine adsorbed decreased as the temperature was raised from 0 to 50 °C. Huang et al. (1984) did not observe any effect of temperature on the adsorption of atrazine by whole soils over a temperature range of 5 to 25°C, while Lavy (1970) reported that diffusion coefficients of many triazines decreased when temperature was lowered from 25 to 5°C, coincident with decreased adsorption. The effect of temperature on the sorption of atrazine on humic acid was quite marked; sorption was nearly twice as great 11 at 40 °C than at 0.5 °C (McGlamery and Slife, 1966). This endothermicity would imply that sorption of atrazine to soil organic matter is governed by favorable entropy changes. Dao and Lavy (1978) reported that greater amounts of atrazine were adsorbed at 30°C than at 5 °C on four soils at 0.1 bar moisture content. They also reported that a decrease in water: soil ratio and in soil moisture content led to an increased adsorption of atrazine. Thus, the influence of temperature on the adsorption by soils seems to be quite variable and it seems to vary with the nature of the soil components (Harris and Warren, 1964). The available data imply that sorption of atrazine to clay minerals is enthalpically favored, while sorption to organic matter is entropically favored. f. Fe and Al oxides Huang et al. (1984) showed that oxides of iron and aluminum play an important role in the adsorption of atrazine by soils and that their removal drastically decreased the adsorption of atrazine by the soils. Terce and Calvet (1977) also reported similar results, while Seta and Karathanasis (1997), and Laird et al. (1994) reported that high contents of iron- and aluminum- oxyhydroxides suppressed atrazine adsorption in soils. This suppression could be due to the inhibition of smectite swelling by these oxyhydroxides. Laird et al. (1994) also observed that these free iron- and aluminum- oxyhydroxides coat silicate minerals, rendering their surfaces more hydrophilic and thus suppressing sorption of hydrophobic compounds like atrazine. Atrazine adsorption-desorption is the major process affecting atrazine behavior in soils and is affected by various factors such as organic matter, clay content, pH, temperature, and iron- and aluminum- oxyhydroxides. We can see that there are many 12 discrepancies in the literature on the effect of inorganic fractions of soils, so atrazine adsorption to minerals needs to be further studied with new methodologies. Molecular Simulations are ideally suited to contribute to our understanding of pesticide behavior in soils because they simultaneously provide mechanistic insight (that can be compared with the available spectroscopic information) and kinetic and thermodynamic estimates (that can be correlated with the many bulk measurements of pesticide diffusion, adsorption isotherms, and distribution coefficients). Thus, we propose to study the atrazine interactions with smectites using molecular simulation methods. Proposal A. Simulating atrazine-clay interactions To simulate the adsorption of atrazine by clays, several types of computational studies will be performed. The minimum-energy structure for atrazine will be computed using density functional theory (Kohn et al., 1996) as implemented in the ab initio programs PQS and Gaussian98 (Frisch et al., 1998). The B3LYP functionals (Becke, 1993) were used with the 6-311G** basis set (Krishnan et al., 1980). For the optimized structure, partial charges for each atom will be estimated using the CHELPG method (Breneman and Wiberg, 1990). Models for K-SWy-2 and K-SAz-l will be created and molecular dynamics simulations of atrazine-clay complexes will be performed in order to explore the interactions between atrazine, cations and water in the interlayer region. The model for K-SWy-2 will have composition of K7(A142Mg6)(Si%Al1)Oz40(OH)43 with 3 CBC of 79 cmol kg'l, while the model for K-SAz-l will have composition of K12(Al3ng.2)Si%0240(OH)43 with a CEC of 132 cmol kg". Atrazine will be added to each unit cell, with the atrazine molecular plane perpendicular to the basal planes of the clay. The interactions between atrazine, a variety of interlayer cations, and water in the interlayer region will be observed at different loading rates, that is, different numbers of atrazine molecules will be added to the unit cell. Water molecules will also be added (in random interlayer positions) so that constant-pressure simulations result in clay layer spacings corresponding to 1, 2, or 3 discrete layers of interlayer water. Target dom— spacings for these simulations will be about 12.5, 15.5, and 18.5 A, respectively. The energies will be computed using a force field deveIOped especially for clays (Teppen et al., 1997) combined with the pcff force field (Maple et al., 1994; MSI, 2000) for the organics and water. The molecular dynamics simulation will be run in the NPT ensemble for 0.1 ns, much longer than the time required for the system volume and energy to equilibrate. 8. Experimental studies of atrazine-clay interactions Adsorption of atrazine on smectites has been studied by a number of workers, but most of these studies have been limited to montmorillonite, beidellite, and nontronite, all of which are 2:1 dioctahedral clays. Only a limited number of studies have been conducted with 2:1 trioctahedral clays. Also, there have been very few studies that systematically investigated atrazine sorption to smectites as a function of the saturating cation, as has been shown to be so important in the sorption of other neutral organics (Sheng et al., 2002). These limited studies indicate that 2:1 trioctahedral clays Often have higher adsorption capacities for atrazine than 2:1 dioctahedral clays. For example, Laird l4 et al. (1992) worked with 14 different clays and found that beidellite, hectorite, and saponite sorbed almost all of the added atrazine. We propose to conduct a laboratory experiment that will study the adsorption of atrazine by saponite, a 2:1 trioctahedral clay, saturated with Na+, K+, Cs+, and Ca2+ on the exchange complex. Depending on the results, we may extend this study to include other triazines: Simazine and metribuzin are two other pesticides belonging to the class of triazines, but with very different water solubilities than atrazine. Simazine is only slightly soluble in water (5 mg/L), while metribuzin has a solubility of 1200 mg/L. Due to differences in water solubility, these triazines are expected to sorb differently in soils. If the saponite results are interesting, we may also use another clay mineral such as hectorite or beidellite to study the adsorption of these triazines: Both hectorite (a trioctahedral smectite like saponite, but differing in that hectorite has dominantly octahedral substitution) and beidellite (dominated by tetrahedral substitution, like saponite, but differing in that beidellite is dioctahedral) showed strong affinity for atrazine in Ca-saturated systems (Laird et al., 1992). C. Prediction of organic-clay interactions Many of the organic pollutants in the enviromnent were introduced years or sometimes decades ago at a time when industry and the public were not adequately aware of the scope, magnitude, and importance of environmental pollution (Alexander, 2000). Thousands of chemicals are manufactured world-over and many new chemicals are added to this ever expanding list. Many of these chemicals end up in the environment (e. g. soil, groundwater, and surface water etc.), where they are a possible source of 15 concern. Adsorption and transport of many of these chemicals, especially pesticides, have been studied for more than 40 years. However, it would require unfeasible financial and scientific resources to study the partition constants of such an immense range of chemicals with an equally high number of different natural phases. Thus there is interest in predictive models that can estimate the potential for a given organic compound to partition into a given sorptive phase (G035 and Schwarzenbach, 2001). We propose to use molecular dynamic simulations to estimate the adsorption energies of some chemicals that have been heavily studied from an experimental standpoint, and then correlate these adsorption energies with the adsorption coefficients obtained from the experiments. If the computed adsorption energies follow the experimental trends in adsorption coefficients, then we will have demonstrated the potential for molecular simulations to predict the sorption of neutral organic compounds to clay minerals. To estimate the adsorption energies, we will simulate a series of nitroaromatic compounds interacting with smectite clay minerals as studied in a recent experimental dataset (Boyd et al., 2001). This experimental study is very useful for validation purposes because it reports sorption of 14 nitrobenzene compounds, differing only in the identity or location of a second functional group. A model for the K-saturated SAz-l smectite clay mineral used in the experimental study will be created, with composition K12(Al36Mg12)Si960240(OH)48 and a resultant cation-exchange capacity (C EC ) of 132 cmol/kg. In order to prepare for molecular dynamics Simulations of nitrobenzene- clay complexes, the interlayer region of the clay will be expanded to 20 A and one molecule from a series of substituted nitrobenzenes will be added to the unit cell. with the nitrobenzene molecular plane perpendicular to the basal planes of the clay. Water 16 molecules will then be added (in random interlayer positions) and molecular dynamics performed on each clay-water-organic system. The energies and forces will be computed using a force field developed especially for clays (Teppen et al., 1997) and combined with the pcff force field (Maple et al., 1994; M81, 2000) for the organics and water. liach molecular dynamics Simulations will be run in the NPT ensemble for at least 50 ps. which is about twice as long as required for the system volume and energy to equilibrate. If the initial Simulation does not result in a clay layer Spacing of 12.3 A. then water molecules will either be removed or added (after expanding the clay) and the simulation re-run so that the equilibrium clay layer Spacing is 12.3 i 0.1 A, in accord with the experimental data for substituted nitrobenzenes in air-dried films of K-SAz-l (Sheng et al.. 2001; Sheng et al., 2002). A trajectory of each simulation, containing equally spaced snapshots of geometries, energies, and velocities, will be recorded for later analysis. We will use the following algorithm to calculate the clay-organic interaction energy. E=E m1 ,0, — El — E2; where °E,O, = Total clay—organic system energy 0E] = Energy of system after organic has been deleted °E2 = Energy of the system after clay and water have been deleted °Compute Em, for at least 150 snapshots and average Electrostatic and van der Waal components Of Eml will be calculated the same way, in order to estimate the contributions of each component to the overall sorption energy. This clay-organic interaction energy (Em) is only a portion of the total adsorption energy, but in the following we also propose a method for using molecular Simulations to estimate the overall adsorption energy. The typical experimental scheme for measuring total adsorption energy is Claytaq) + orgm) t::> clay-orgtaq) + HZOtaql Eexpt where Eexp, is the experimentally determined adsorption energy (for example, the overall free energy, enthalpy, or entropy of the transfer process). G055 and Schwarzenbach (2001) illustrated that the overall free energy change of transfer AG, , could be first decomposed into individual terms for a more mechanistic understanding of the partition process as Shown in Figure 1. l8 Partitioning of small organics between a clay phase and an aqueous phase Clay phase Aqueous phase AG.- = E. + Bitty mystic?“ ’3‘” Figure 1. Schematic for decomposing the energy changes for transfer of a compound from aqueous solution to the clay layer (after Goss and Schwarzenbach, 2001). Since the experimental reaction is too large and too slow to directly model with molecular simulation methods, an alternative path is proposed that is more computationally feasible and follows the spirit of Figure 1. We propose to employ the following cycle for computing the total adsorption energy: Claym) + orgtaq) => clay-orguq) + H200”, Eexpt Clay-cavity(,q) + orgm) + H20 => clay-organ) + cavitym) + H20 E4 This approach is similar to many schemes (i.e., see review in (Halperin et al., 2002)) that have been used to compare molecular simulations with experimental data for the binding of drugs to polypeptides. Since the free energy, enthalpy, and entropy are state functions (Atkins, 1998), their magnitudes should be independent of path, so that Eexpt : E3 + E4 + E5 WhCI‘C E4 = Ei-clay—Ei-aqueous = E(clay-cav —> clay-org) + E(orglaq) ——->0rgca\'(aq)) E4 = Em: + E(org(aq> —> ca\'ily(aq)) E3 is the energy for creating a cavity in the clay, by removing enough water molecules to make a cavity that can later receive the organic. E3 will be calculated in the same way as the algorithm above for Em, using the smectite with a 12.3 A d-spacing and only cations and water in the interlayer. E4 is similar to the sum of two components as shown in Figure 1. While we are not using the gas-phase configuration that Goss and Schwarzenbach (2001) used, our term E4 can be estimated the same way. The first term in E4 is Em, the energy for filling the cavity in the clay with the organic molecule. The second term in E4 is the energy for moving the organic solute out of the aqueous phase, thereby leaving an empty cavity in water. E5 is the energy for filling this aqueous cavity with water, and E5 will again be calculated using the algorithm for Eim that was presented above. 20 We will use all of the nitroaromatic compounds from the experimental database (Boyd et al., 2001) and calculate the clay-organic interaction energies for each compound using the algorithm above for Eim. Then we will fit these interaction energy values into the larger context, using the thermodynamic cycle given above to estimate the overall adsorption enthalpies for each compound. The enthalpy estimates will be checked against experimental data in two ways. In relative terms, we hypothesize that the enthalpy estimates will follow the same trends as the sorption coefficients for the nitroaromatic compounds (Boyd et al., 2001). In absolute terms, we hypothesize that our enthalpy estimates will compare favorably with experimental enthalpies recently estimated from the temperature dependence of adsorption isotherms (Li et al., 2004). If predictive capability can be demonstrated by using these nitroaromatic compounds as a training set, then we will use the methods to estimate the enthalpies of adsorption for atrazine in smectite clay minerals. 21 LITERATURE CITED Alexander, M. 2000. Aging, bioavailability, and overestimation of risk from environmental pollutants. Environmental Science & Technology 34:4259-4265. Atkins, P.W. 1998. Physical chemistry. 6th ed. W.H. Freeman and Co., New York, NY. Barriuso, E., D.A. Laird, W.C. Koskinen, and RH. Dowdy. 1994. Atrazine desorption from smectites. Soil Science Society of America Journal 58:1632-1638. Baskaran, 8., NS. Bolan, A. Rahman, and R.W. Tillman. 1996. Pesticide sorption by allophanic and non-allophanic soils of New Zealand. New Zealand J oumal of Agricultural Research 39:297-310. Becke, AD. 1993. A New Mixing of Hartree-Fock and Local Density-Functional Theories. Journal of Chemical Physics 98:1372-1377. Belluck, D.A., S.L. Benjamin, and T. Dawson. 1991. Groundwater contamination by atrazine and its metabolites: risk assessment, policy and legal implications, In L. Somasundaram and J. R. Coats, eds. Pesticide transformation products: fate and significance in the environment. American Chemical Society, Washington, DC. Biradar, DP, and AL. Rayburn. 1995. Chromosomal Damage-Induced By Herbicide Contamination At Concentrations Observed in Public Water-Supplies. J oumal of Environmental Quality 24: 1222-1225. Blanchard, PE, and W.W. Donald. 1997. Herbicide contamination of groundwater beneath claypan soils in North-Central Missouri. J oumal of Environmental Quality 26: 1 612-1621. Borchardt, G. 1989. Smectites, p. 675-727, In J. B. Dixon and S. B. Weed, eds. Minerals in soil environments, 2nd ed. Soil Science Society of America, Madison, WI. Borggaard, OK, and J .C. Streibig. 1988. Atrazine adsorption by some soil samples in relation to their constituents. Acta Agriculturae Scandinavica 38:293-301. Boyd, S.A., G.Y. Sheng, B.J. Teppen, and OJ. Johnston. 2001. Mechanisms for the adsorption of substituted nitrobenzenes by smectite clays. Environmental Science & Technology 35:4227-4234. Breneman, C.M., and KB. Wiberg. 1990. Determining atom-centered monopoles from molecular electrostatic potentials. The need for high sampling density in formamide conformational analysis. Journal of Computational Chemistry 1 1:361- 373. 22 Brouwer, W.W.M., J.J.T.I. Boesten, and W.G. Siegers. 1990. Adsorption of transformation products of atrazine by soil. Weed Research 30: 123-128. Brown, G., A.C.D. Newman, J .H. Rayner, and AH. Weir. 1978. The structure and chemistry of soil clay minerals, p. 29-178, In D. J. Greenland and M. H. B. Hayes, eds. The chemistry of soil constituents. John Wiley and Sons, Chichester, UK. Burkart, M.R., D.W. Kolpin, and DE. James. 1999. Assessing groundwater vulnerability to agrichemical contamination in the Midwest. Water Science and Technology 391103-112. Chiou, C.T., L]. Peters, and V.H. Freed. 1979. Physical concept of soil-water equilibria for nonionic organic compounds. Science 206:831-832. Chiou, C.T., P.E. Porter, and D.W. Schmedding. 1983. Partition Equilibria ofNon-lonic Organic-Compounds Between Soil Organic-Matter and Water. Environmental Science & Technology 17:227-231. Chiou, C.T., T.D. Shoup, and RE. Porter. 1985. Mechanistic roles of soil humus and minerals in the sorption of nonionic organic-compounds from aqueous and organic solutions. Organic Geochemistry 8:9-14. Chiou, C.T., D.E. Kile, D.W. Rutherford, G.Y. Sheng, and SA. Boyd. 2000. Sorption of selected organic compounds from water to a peat soil and its humic-acid and humin fractions: Potential sources of the sorption nonlinearity. Environmental Science & Technology 34:1254-1258. Clay, S.A., W.C. Koskinen, R.R. Allmaras, and RH. Dowdy. 1988. Differences in herbicide adsorption on soil using several several soil pH modification techniques. Journal of Environmental Science and Health B 23:559-573. Dao, TH, and TL. Lavy. 1978. Atrazine adsorption on soil as influenced by temperature, moisture content and electrolyte concentration. Weed Science 26:303-308. Frisch, M.J., G.W. Trucks, H.B. Schlegel, G.E. Scuseria, M.A. Robb, J .R. Cheeseman, V.G. Zakrzewski, J .A. Montgomery, R.E. Stratmann, J .C. Burant, S. Dapprich, J.M. Millarn, A.D. Daniels, K.N. Kudin, M.C. Strain, O. Farkas, J. Tomasi, V. Barone, M. Cossi, R. Cammi, B. Mennucci, C. Pomelli, C. Adamo, S. Clifford, J. Ochterski, G.A. Petersson, P.Y. Ayala, Q. Cui, K. Morokuma, D.K. Malick, A.D. Rabuck, K. Raghavachari, J .B. Foresman, J. Cioslowski, J .V. Ortiz, B.B. Stefanov, G. Liu, A. Liashenko, P. Piskorz, I. Komaromi, R. Gomperts, R.L. Martin, D.J. Fox, T. Keith, M.A. Al-Laham, C.Y. Peng, A. Nanayakkara, C. Gonzalez, M. Challacombe, P.M.W. Gill, B.G.Johnson, W. Chen, M.W. Wong, J .L. Andres, M. Head-Gordon, R.E. S., and J .A. Pople. 1998. Gaussian 98, Revision A.9 Gaussian, Inc., Pittsburgh, PA. 23 Frissel, M.J., and CH. Bolt. 1962. Interaction between certain ionizable organic compounds (herbicides) and clay minerals. Soil Science 94:284-291. Gilchrist, G.F.R., D.S. Gamble, H. Kodama, and S.U. Khan. 1993. Atrazine interactions with clay-minerals - kinetics and equilibria of sorption. Journal of Agricultural and Food Chemistry 41:1748-1755. Gilmour, J .T., and NT. Coleman. 1971. s-Triazine adsorption studies: Ca-H-Humic acid. Proceedings of the Soil Science Society of America 35:256-259. Goss, K.U., and RP. Schwarzenbach. 2001. Linear free energy relationships used to evaluate equilibrium partitioning of organic compounds. Environmental Science & Technology 3521-9. Grathwohl, P. 1990. Influence of Organic-Matter From Soils and Sediments From Various Origins On the Sorption of Some Chlorinated Aliphatic- Hydrocarbons - Implications On Koc Correlations. Environmental Science & Technology 24:1687-1693. Grundl, T., and G. Small. 1993. Mineral contributions to atrazine and alachlor sorption in soil mixtures of variable organic carbon and clay content. Journal of Contaminant Hydrology 14:117-128. Haderlein, 8.8., and RP. Schwarzenbach. 1993. Adsorption of substituted nitrobenzenes and nitrophenols to mineral surfaces. Environmental Science & Technology 27:316-326. Haderlein, S.B., K.W. Weissmahr, and RP. Schwarzenbach. 1996. Specific adsorption of nitroaromatic explosives and pesticides to clay minerals. Environmental Science & Technology 30:612-622. Halperin, 1., BY. Ma, H. Wolfson, and R. Nussinov. 2002. Principles of docking: An overview of search algorithms and a guide to scoring functions. Proteins-Structure Function and Genetics 47:409-443. Hamaker, J .W., and J.M. Thompson. 1972. Adsorption, p. 49-143, In C. A. I. Goring and J. W. Hamaker, eds. Organic chemicals in the soil environment, Vol. 1. Marcel Dekker, New York. Hance, R]. 1965. Observations on the relationship between the adsorption of diuron and the nature of the adsorbent. Weed Research 5: 108-1 14. Harris, CL, and GR Warren. 1964. Adsorption and desorption of herbicides by soil. Weeds 12:120-126. Hayes, M.H.B., and U. Mingelgrin. 1991. Interactions between small organic chemicals and soil colloidal constituents, p. 323-407, In G. H. Bolt, ed. Interactions at the soil colloid-soil solution interface. Kluwer Academic Publisher, Boston. 24 Huang, P.M., R. Grover, and RB. McKercher. 1984. Components and particle-size fractions involved in atrazine adsorption by soils. Soil Science 138:20-24. Huang, W.L., T.M. Young, M.A. Schlautman, H. Yu, and W.J. Weber. 1997. A distributed reactivity model for sorption by soils and sediments .9. General isotherm nonlinearity and applicability of the dual reactive domain model. Environmental Science & Technology 31 :1703-1710. Karickhoff, SW. 1984. Organic pollutant sorption in aquatic systems. Journal of Hydraulic Engineering 1102707-735. Khan, A., J .J . Hassett, W.L. Banwart, J .C. Means, and SC. Wood. 1979. Sorption of acetophenone by sediments and soils. Soil Science 128:297-302. Kleineidam, S., H. Rugner, B. Ligouis, and P. Grathwohl. 1999. Organic matter facies and equilibrium sorption of phenanthrene. Environmental Science & Technology 33:1637-1644. Kohn, W., A.D. Becke, and RC. Parr. 1996. Density functional theory of electronic structure. Journal of Physical Chemistry 100: 12974-12980. Kolpin, D.W., E.M. Thurman, and DA. Goolsby. 1996. Occurence of selected pesticides and their metabolites in near-surface aquifers of the midwestem United States. Environmental Science & Technology 30:335-340. Kolpin, D.W., J.B. Barbash, and R]. Gilliom. 1998. Occurrence of pesticides in shallow groundwater of the United States: Initial results from the National Water-Quality Assessment Program. Environmental Science & Technology 32:558-566. Krishnan, R., J .S. Binkley, R. Seeger, and J .A. Pople. 1980. Self-consistent molecular orbital methods. XX. A basis set for correlated wave functions. The Journal of Chemical Physics 72:650-654. Laird, DA. 1996. Interactions between atrazine and smectite surfaces, p. 86-100, In M. T. Meyer and E. M. Thurman, eds. Herbicide Metabolites in Surface Water and Groundwater, Vol. 630. American Chemical Society, Washington, DC. Laird, D.A., E. Barriuso, R.H. Dowdy, and W.C. Koskinen. 1992. Adsorption ofatrazine on smectites. Soil Science Society of America Journal 56:62-67. Laird, D.A., P.Y. Yen, W.C. Koskinen, T.R. Steinheimer, and RH. Dowdy. 1994. Sorption of atrazine on soil clay components. Environmental Science & Technology 28: 1054- 1 061. Lavy, TL. 1970. Diffusion of three chloro s-triazines in soil. Weed Science 18:53-56. 25 Li, H., B.J. Teppen, C.T. Johnston, and SA. Boyd. 2004. Thermodynamics of nitroaromatic compound adsorption from water by smectite clay. Environmental Science & Technology 38:5433-5442. Macnamara, G., and SJ. Toth. 1970. Adsorption of linuron and malathion by soils and clay minerals. Soil Science 109:234-240. Madsen, L., B. Lindhardt, P. Rosenberg, L. Clausen, and I. Fabricius. 2000. Pesticide sorption by low organic carbon sediments: a screening for seven herbicides. Journal of Environmental Quality 29: 1488-1 500. Maple, J.R., M.-J. Hwang, T.P. Stockfish, U. Dinur, M. Waldman, C.S. Ewig, and AT. Hagler. 1994. Derivation of Class H force fields. 1. Methodology and quantum force field for the alkyl functional group and alkane molecules. Journal of Computational Chemistry 15:162-182. McBride, MB. 1994. Environmental chemistry of soils Oxford University Press, New York. McGinley, P.M., L.E. Katz, and W.J. Weber. 1993. A distributed reactivity model for sorption by soils and sediments. 2. Multicomponent systems and competitive effects. Environmental Science & Technology 27: 1524-1531. McGlamery, M.D., and F.W. Slife. 1966. The adsorption and desorption of atrazine as affected by pH, temperature, and concentration. Weeds 14:237-239. Means, J .C., S]. Wood, J .J . Hassett, and W.L. Banwart. 1982. Sorption of amino- and carboxy-substituted polynuclear aromatic hydrocarbons by sediments and soils. Environmental Science & Technology 16:93-98. Accelrys, Inc. 2000. Cerius2 Molecular Simulation Software, Release 4.2. Accelrys, Inc., San Diego, CA. Pike, DR, and EL. Knake. 1996. Field corn and soybean pesticide use summaries. 1991-1994 National agricultural statistical service data on specific products. Special Report 1996-02. Department of Crop Sciences, University of Illinois, Urbana-Champaign. Rebhun, M., R. Kalabo, L. Grossman, J. Manka, and C. Ravacha. 1992. Sorption of organics on clay and synthetic humic clay complexes simulating aquifer processes. Water Research 26:79-84. Ribaudo, M.O., and A. Bouzaher. 1994. Atrazine: Environmental characteristics and economics of management. microfonn Al .1072699. US. Department of Agriculture, Economic Research Service, Herndon, VAzERS-NASS, Washington, DC. 26 Roy, W.R., and LG. Krapac. 1994. Adsorption and desorption of atrazine, deethylatrazine by low organic carbon geologic material. Journal of Environmental Quality 23:549-556. Sawhney, BL, and SS. Singh. 1997. Sorption of atrazine by Al- and C a-saturated smectite. Clays and Clay Minerals 45:333-338. Seta, AK, and AD. Karathanasis. 1997. Atrazine adsorption by soil colloids and C 0- transport through subsurface environments. Soil Science Society of America Journal 61:612-617. Sheng, G.Y., C.T. Johnston, B.J. Teppen, and SA. Boyd. 2001. Potential contributions of smectite clays and organic matter to pesticide retention in soils. Journal of Agricultural and Food Chemistry 49:2899-2907. Sheng, G.Y., C.T. Johnston, B.J. Teppen, and SA. Boyd. 2002. Adsorption of dinitrophenol herbicides from water by montmorillonites. Clays and Clay Minerals 50:25—34. Talbert, R.E., and OH. Fletchall. 1965. The adsorption of some s-triazines in soils. Weeds 13:46-52. Teppen, B.J., K. Rasmussen, P.M. Bertsch, D.M. Miller, and L. Schafer. 1997. Molecular dynamics modeling of clay minerals .1. Gibbsite, kaolinite, pyrophyllite, and beidellite. Journal of Physical Chemistry B 101:1579-1587. Terce, M., and R. Calvet. 1977. Some observations on the role of A1 and Fe and their hydroxides in the adsorption of herbicides by montmorillonite. Z. Pflkrankh. Pflschutz Sonderheft 8:237-243. US. Congress Office of Technology Assessment 1984. Protecting the nation’s groundwater from contamination. Volume 2, OTA-O-276, US. Government Printing Office, Washington, DC. Vaz, C.M.P., S. Crestana, S.A.S. Machado, L.H. Mazo, and LA. Avaca. 1997. Adsorption isotherms for atrazine on soils measured by differential pulse polarography. Electroanalysis 92956-958. Wang, Z.D., D.S. Gamble, and CH. Langford. 1990. Interaction Of atrazine with Laurentian fulvic-acid- Binding and hydrolysis. Analytica Chimica Acta 232: 1 8 l - 188. Weber, J.B. 1970. Mechanisms of adsorption of s-triazines by clay colloids and factors affecting plant availability. Residue Reviews 32293-130. 27 Weber, W.J., and W.L. Huang. 1996. A distributed reactivity model for sorption by soils and sediments .4. Intraparticle heterogeneity and phase- distribution relationships under nonequilibrium conditions. Environmental Science & Technology 30:881- 888. Weber, W.J., P.M. McGinley, and LE. Katz. 1992. A distributed reactivity model for sorption by soils and sediments. 1. Conceptual basis and equilibrium assessments. Environmental Science & Technology 2621955-1962. Welhouse, G.J., and W.F. Bleam. 1992. NMR spectroscopic investigation of hydrogen- bonding in atrazine. Environmental Science & Technology 26:959-964. Welhouse, G.J., and W.F. Bleam. 1993a. Atrazine hydrogen-bonding potentials. Environmental Science & Technology 27:494-500. Welhouse, G.J., and W.F. Bleam. 1993b. Cooperative hydrogen-bonding of atrazine. Environmental Science & Technology 27:500-505. Xia, GS, and WP. Ball. 2000. Polanyi-based models for the competitive sorption of low- polarity organic contaminants on a natural sorbent. Environmental Science & Technology 34:1246-1253. Xing, BS, and J .J . Pignatello. 1997. Dual-mode sorption of low-polarity compounds in glassy poly(vinyl chloride) and soil organic matter. Environmental Science & Technology 31:792-799. Yaron, B., and S. Saltzman. 1972. Influence of water and temperature on adsorption of parathion by soils. Soil Science Society of America Proceedings 36:583-586. Young, T.M., and W.J. Weber. 1995. A distributed reactivity model for sorption by soils and sediments. 3. Effects of diagenetic processes on sorption energetics. Environmental Science & Technology 29:92-97. 28 CHAPTER II TRIAZINE ADSORPTION BY SAPONITE AND BEIDELLITE CLAY MINERALS 29 Triazine adsorption by saponite and beidellite clay minerals Abstract We investigated the sorption of three triazine herbicides (atrazine, simazine, and metribuzin) by saponite and beidellite clay minerals saturated with K+, Cs+, Na+, and Ca2+. Saponite clay sorbed a larger fraction of each pesticide from aqueous solution than did beidellite clay. The lower cation exchange capacity in saponite (vs. beidellite) presumably results in a less crowded interlayer with more siloxane surface available for adsorption. Generally, Cs-saturated clays sorbed more triazines than clays saturated by K+, Na+ or Ca2+. We attribute this to the smaller hydrated radius of C 8+, which again increases the siloxane surface available for sorption. Furthermore, the relatively weak hydration of Cs+ reduces the swelling of clay interlayers, thus making sorption domains less hydrated and more receptive to hydrophobic molecules. C S-saponite manifested a sorption of more than 1% atrazine by weight above equilibrium concentrations of 6 mg/L. This is the largest sorption of neutral atrazine from water yet reported for an inorganic sorbent. Molecular dynamics Simulations indicate that atrazine interacts with both clay basal planes and with multiple cations in the clay interlayer. 30 Introduction Concern about the contamination of ground and surface waters by pesticides has been an area of concern for over thirty years, mainly because of the continuous use of such chemicals in agricultural production. There is a need to more accurately predict the fate and movement of pesticides and assess their risks in soil and subsurface environments. Sorption by soils/sediments is a major process that controls the transport, bioavailability, degradation and ultimately the fate of organic contaminants in the environment (Koskinen and Harper, 1990). Sorption of organic chemicals diminishes their bioavailability to target organisms (Albanis et al., 1988), affects both chemical and microbiological transformations (Armstrong and Chesters, 1968; Gamble and Khan, 1985), and reduces the leaching to groundwater (Sonon and Schwab, 1995). Clay minerals and soil organic matter (SOM) are typically the two most chemically active components interacting with organic chemicals that enter soils and sediments. SOM has been accepted as the primary sorptive domain for the uptake of hydrophobic organic compounds (Chiou et al., 1979; Chiou et al., 2000; Huang et al., 1997; Kleineidam et al., 1999; Weber and Huang, 1996; Xia and Ball, 2000). The contaminant transport models often use soil-organic carbon-normalized sorption coefficients (Koc) to predict the distribution of organic chemicals between soil and water. This approach ignores the contribution of soil minerals to the retention of organic contaminants in soils and subsoils, based on the assumption that organic compounds cannot effectively compete with water molecules for the sorptive sites on mineral surfaces. 31 Clay minerals are commonly found in the soil environment, and smectites are the most important group for interacting with pesticides due to their high surface areas, moderate cation exchange capacities (CECS), and surface reactivities (McBride, 1994). Earlier studies have shown that smectite clays can adsorb many major types of pesticides such as carbamates, ureas, nitrophenols and triazines to an appreciable extent (Li et al., 2004; Li et al., 2003; Sheng et al., 2001). In a study of atrazine sorption by a series of Ca-smectite clays with CECs ranging from 79 to 134 cmolc/kg, Laird et al. (1992) reported that zero to nearly 100% of added atrazine was removed from water, with the extent of adsorption correlated inversely with the clay CEC values. Sheng et al. (2001) observed a similar relationship for sorption of atrazine by K-smectites. Sawhney and Singh (1997) reported that atrazine adsorption by Al-smectite (>1 g kg!) was larger than that by Ca-smectite. The authors discounted protonation of atrazine and attributed the greater adsorption on Al-clay to the stronger H-bonding between atrazine and the more polarized water molecules associated with trivalent Al3+ versus divalent Ca2+. 1n the present study, we have avoided Al-Smectites due to the difficulty in Simultaneously maintaining a homoionic Al-smectite while preventing side-reactions in which atrazine protonates and consequently adsorbs strongly by cation exchange. Clays present in whole soils also play a major role in adsorbing triazine pesticides (Huang et al., 1984; Talbert and Fletchall, 1965). Talbert and F letchall (1965) Showed that atrazine sorption by soils was positively correlated with soil clay content as well as SOM content. Huang et al. (1984) reported that the inorganic fractions of soils with particle size <20 um have a significant capacity to retain atrazine. Several previous studies have attempted to identify the critical ratio of clay minerals/SOM at which 32 sorption by the mineral phase plays an important role (Grundl and Small, 1993; Karickhoff, 1984; Means et al., 1982). Karickhoff (1984) concluded that at the ratio of swelling clays to soil organic carbon >30 the mineral phase in whole soils contributed significantly to sorption of simazine. Grundl and Small (1993) estimated that at a mass ratio near 62 for clay to organic carbon, the mineral phase accounted for half of the overall atrazine sorption. Most studies of atrazine and simazine adsorption by smectites were conducted using smectites with negative charges originating from octahedral substitution (Gilchrist et al., 1993; Sannino et al., 1999; Sheng et al., 2001); few studies have used smectites with negative charges arising from tetrahedral sites. Also, investigations of the effect of the location of permanent negative charges on organic sorption by smectites are scarce. Compared with octahedrally substituted smectites, the negative charges in tetrahedrally substituted smectites are more localized to fewer oxygens on siloxane Sheets (Farmer, 1978; McBride, 1989), resulting in differences in clay interlayer environments that may influence pesticide sorption. Beidellite and saponite clays are smectites with negative charges originating largely from tetrahedral substitution. This study was, therefore, initiated to study the adsorption of three s-triazine herbicides (atrazine, simazine, and metribuzin) by saponite and beidellite clays saturated with different cations (KI, Na+, C 5+ and Ca2+), and to compare the results with previous studies on sorption by octahedrally substituted smectites. K+, Na+ and Ca2+ are common exchangeable cations in the soil environment, while Cs+ has smaller hydrated radius relative to other cations studied here hence is expected to create a more favorable (i.e., less hydrated) clay interlayer environment for pesticide adsorption. 33 Materials and Methods Clays and Herbicides Two reference clays, saponite (SapCa-Z, CEC 95 cmol/kg (Post, 1984)) and beidellite (SBId-l, CEC 129 cmol/kg (Hetzel and Doner, 1993)), were obtained from the Clay Minerals Society Source Clay Repository (Department of Agronomy, Purdue University, West Lafayette, Indiana, USA). The <2 urn clay-sized particles were separated by wet sedimentation, and then exchanged with K+, Na+, CS+, and C a2+ to prepare homoionic clays. To do so, 25 g of clay was dispersed in a 0.1 M chloride solution (500 ml) with the respective cation. The clay suspensions were shaken for 24 h, and fresh chloride salt solutions were used to replace the original solutions after centrifugation. This process was repeated four times to ensure complete cation saturation. The excess chlorides were removed by repeatedly washing with Milli-Q water until Cl’ was determined negative by reacting with AgNO; solution. The clay suspensions were then quick-frozen, freeze-dried and stored in closed containers for later use. The pH values of the smectite suspensions were 8.1 and 8.0 for Ca-saponite and -beidellite, respectively, 8.6 and 8.8 for the Na-smectites, 8.1 and 8.0 for the C s- smectites, and 8.2 and 8.0 for the K-smectites. The three pesticides (atrazine, simazine, and metribuzin) used in the adsorption experiment were purchased from ChemService (West Chestnut, Pennsylvania) with purities of >99%, and used as received. The selected physical and chemical properties of the three pesticides along with their chemical structures are summarized in Table 1. 34 Z Axum luv ,2 _ __ . . . z/ \oluxfuv owm co _ cm? 0 _ om Em xw ENBEDE O fmozz iwszxwt zzfmc z z \ z 5 Womm omm w mo." 8. SN JO MEMBER 5 x z / m:Nozz AK wizzfmu 2,, \z omm SN mm on; 09m 5. w “SENSE 5 A85 A695 EwEB Sang Sev— w2 8me E 33 836222 2:88am gsaem Siam zeta—com £5 5 com: moEocmom we wear—2.8m _ao_Eo:ooooo_m.Em tote—om A 03:. Given the pKa values of the herbicides and the high pH values of our clay suspensions,there was no possibility for triazine sorption as any cationic species in the present study. Rather, the triazines were all neutral, moderately hydrophobic solutes. Sorption Isotherm Measurements Pesticide sorption isotherms were determined for each clay using a batch equilibration method. A series of initial pesticide concentrations was prepared by dissolving each pesticide in 0.01 M KCl, NaCl, and CSCl, and 0.005 M CaClz. The initial solutions were then added into glass centrifuge tubes containing clay (0.01 to 0.1 g) that was saturated with the same cation in the respective pesticide initial solution, and the tubes were closed with Teflon-lined screw caps. Duplicate samples were prepared to measure sorption. The ratio of solution volume to clay amount was optimized to make sure that 20-80 % of pesticide in solution phase was sorbed by clay during the equilibration. The tubes were then Shaken reciprocally for 12 hours at 40 rpm at room temperature (23 i 2 °C); preliminary studies showed sorption was complete within this time. The tubes were then centrifuged at 2110 g for 20 min to separate the liquid and solid phases. The concentration of herbicide in supernatant was analyzed using a reversed-phase high-performance liquid chromatography (HPLC, Perkin-Elmer, Norwalk, CT) set at the appropriate wavelength for maximum absorption by each pesticide (Table 1). A platinum extended polar selectivity (EPS) C18 column was used. The mobile phase was a mixture of either 65% methanol (for atrazine and metribuzin) or 65% acetonitrile (for simazine) and water with a flow rate of 1.0 mL/min. Controls consisted of the initial pesticide solutions in the supporting electrolyte. Abiotic 36 degradation of atrazine and Simazine in water at pH 7 to 9 is negligible (Comber, 1999; Noblet et al., 1996) in the absence of strong UV light. However, due to the relatively high pH values in the clay slurries and the lack of information about possible smectite- catalyzed hydrolysis, we monitored the HPLC chromatograms for peaks indicative of triazine degradation products. None were observed, and there were no significant differences between peaks or shapes of chromatograms for triazine controls versus triazine in clay suspensions. Also, no changes in solute concentrations were detected in the tubes devoid of clay within the experimental period; therefore solute mass lost in the supernatant from clay slurries was assumed to be sorbed by the clay. The amount of pesticide sorbed on the clays was calculated from the difference between the amount added and that remaining in the final equilibrated solution. Data presented are the means of duplicate samples. X-ray Diffraction After supematants were sampled for HPLC analysis, the remaining pesticide-clay suspensions were used to prepare clay films for X-ray diffraction analysis. The clay solids were resuspended with the supernatant (~ 1 to 2 mL) remaining in the vial, and then dropped onto glass slides and air-dried overnight at ambient conditions to Obtain oriented clay films. X-ray diffraction patterns were recorded using Cu-KOL radiation and a Philips APD 3720 automated X-ray diffractiometer, fitted with an APD 3521 goniometer and a 0-compensating slit, a 0.2 mm receiving slit and a diffracted beam graphite monochromator. Diffraction patterns were recorded from 5.5 to 16.0 °20 at a step of 0.005 °20 with 4 s/step. The X-ray diffraction patterns were re-recorded after the 37 clay films were allowed to equilibrate with 100% relative humidity for 2 d in a closed container. Molecular Simulations Molecular dynamics simulations of atrazine-clay complexation were performed in order to explore possible interaction mechanism(s) between atrazine, cations, and water in the clay interlayer regions. In this approach, electrons are not treated explicitly, but their effects are parameterized into classical (that is, non-quantum) mechanical energy functions comprising a force field. The model for K-saponite had composition K9(Sig7A19)Mg720240(OH)48 with a CEC of 95 cmol/kg. The smectite surfaces were constructed using methods previously detailed (Teppen et al., 1997), in which a muscovite mica layer was the starting point and Al-for-Si substitutions were random except that no two A1 tetrahedral were allowed to be immediate neighbors. One molecule of atrazine was added to the unit cell, with the atrazine molecular plane perpendicular to the basal planes of the clay. The simulated loading rate was equivalent to 23 mg atrazine per g clay. Water molecules were also added (in random interlayer positions) so that constant-pressure simulations resulted in ~12.5 A clay layer spacings, in accord with our experimental data for atrazine in films of K-saponite. The energies were computed using a force field developed for clays (Teppen et al., 1997) combined with the polymer consistent force field (PCFF) (Maple et al., 1994; M81, 2000) for the organics and water. AS a further modification, charges on the organic solute were assigned using the COMPASS force field (Sun, 1998), which optimized organic atomic charges for condensed phase systems. In previous work (Boyd et al., 38 2001; Sheng et al., 2002), we have found COMPASS charges to more faithfully reproduce quantum mechanical charges for nitrogenous organic molecules than do PC F F charges. In aqueous interfacial systems, the charges on atoms of organic solutes play a critical role in determining organic solute interactions with the clay and interlayer water and cations. In classical molecular dynamics, the nuclear charge and the local electron density are subsumed into one partial atomic charge and these charges for atrazine are illustrated in Figure l. ('31 C1 N1/ \\N1 CII (I: Charge H / 2\ ¢ 2‘.‘ H1 C1 +0.5830 ‘\N’,.-" N1 ‘\.\ / C2 +0.5637 7— H N2 C3 +0.1048 2 C4 -0.1590 H 2\c3/ H2 C5 +0.1578 H - 2\C /C4—H2 Nl 0.4810 5 \ N2 -0.5422 C4 H2 H1 +0.2487 2 C4—H2 C1 -01020 Hz /\ H2 Figure 1. Charges on atrazine molecule used in simulation study. 39 Commercial software (M81, 2000) was used to compute the classical molecular dynamics. No constraints were imposed on our systems during their molecular dynamics except for the contents of the simulation cell: All atoms were completely free to move, and unit cell shape and volume were unconstrained. The molecular dynamics simulations were run in the NPT (constant pressure/constant temperature) ensemble for 75 ps, which is much longer than the time required for the system volume and energy to equilibrate. Results and Discussion Sorption Isotherm Pesticide sorption isotherms are shown in Figure 2. The greatest sorption of atrazine was observed for Cs-saturated saponite, followed by K-, Na-, and least by C a- saturated saponite (Figure 2a). This sorption of more than 1% atrazine by weight of clay at an aqueous atrazine concentration of 6 mg/L is by far the largest sorption of neutral atrazine from water yet reported for any inorganic material. The higher effectiveness of Cs-SapCa-2 in sorption may be due to the lower hydration energy of Cs+ and hence smaller hydrated radius of Cs+ compared to the other cations. The water molecules associated with exchangeable Ca2+, for example, obscure a larger fraction of the clay surface than those associated with Cs+, resulting in the reduction in the size of adsorptive domains for Ca2+ saturated clays. Typical estimates of hydrated radii are 9.6 A for C a” and 3.6 A for CS+ (Evangelou, 1998). Thus the cross-sectional area of hydrated C a2+ is about 290 A2, more than seven times that of Cs+(40 A2); therefore, one hydrated Ca2+ takes up 3.5 times as much room as do the two Cs+ ions in clay interlayer. We 40 hypothesize that the CS-smectites may sorb triazines more strongly because the triazines only have to compete with weakly bound water between hydrated cations, while on C a-, K-, and Na-smectites the triazines must compete with more strongly bound water within the hydrated radii of cations. Sheng et al. (2001, 2002) also observed similar effects of exchangeable cations on the Size of sorptive domains of smectite clays in pesticide sorption. 41 16 - - _-. _, I (a) Atrazine Saponite I 14 p ' . 12 1 on- I V 10 1; 0 . ' V 0‘ . 8 ‘1 I 9 . I v 01 o ‘ l 6 I - ‘ 'V o . a 4 I 02 ‘ IV 0‘. 2 I 0009’ - . o 2 4 O I 10 12 u. 0 In—WMJ‘FEL ,_ ' _, 0 2 4 6 8 10 12 14 a 1.4 El (c) Mctribuzin Saponite E 1.2 1 V B y " V 'E 1.0 . 0°: v g 0.8 1 .2 ' E} 0.6 « . G- . . h— o 0.4 1 ‘ ‘ A AA ‘15: v V O . ‘A ~ E 0.2 'v ‘ . I I . < A, . 0.0 05-1.. .4 . . . 0 10 20 3O 40 50 ()0 2.5 ' (e) Simazine Saponite 2.0 1 V V . K v 1.5 1 , A Ca ' I Na v Cs 1.0 1 v' 0.51). . . I I A -. «T 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 0.8 1 0.6 1 0.4 1 0.2 0.0 a. V I (b) Atrazine Betdcllttc I 0.7 0.6 1 0.5 1 0.4 1 0.3 1 0.2 1 0.11 0.0 80 l 00 0.4 0.3 1 0.2 1 0.11 (t) Simazine Beidellite V V O O 0.0 O Equilibrium Concentration of Pesticide (mg/L) Figure 2. Sorption isotherms representing atrazine (a,b), metribuzin (c,d), and simazine (e,i) uptake from water by reference saponite (SapCa-Z) and beidellite (SBId-l) clays saturated with CS+, 1C, Na“ and Ca“. 42 Differences in the swelling behavior between Cs- and Ca-SapCa-2 also affect the degree of pesticide adsorption. The relatively weak hydration of Cs+ causes Cs-saponite to swell less in water, which again provides a more favorable sorption domain and allows the organic compound to better compete with waters for interlayer sites. C s-smectites maintain ~12.3 A basal spacings (monolayer of hydration) at 100% humidity (data not shown) and even in aqueous suspension (Sato et al., 1992; Suquet and Pezerat, 1987). Ca-smectites, on the other hand, swell to more than 15 A basal spacings. The thickness of the ring of an atrazine molecule is ~3.0 A (Sheng et al., 2002). This would imply the lowest limit of the basal spacing for atrazine intercalation to be about 12.2 A (the thickness of a clay layer alone for pyrophyllite is 9.2 A). In this scenario, an atrazine molecule thus contacts both clay siloxane surfaces simultaneously (Sheng et al., 2001 ). This orientation favors the intercalation of organic contaminants and pesticides, because it minimizes the contact with water molecules. The efficacy of such a hydrophobic adsorption mechanism even in inorganic smectites is implied by observations (Li et al., 2004; Sheng et al., 2002) that intercalated organic solutes inhibit swelling when the smectite is exposed to water. Our results for atrazine sorption by saponite clay compare favorably with the results reported by Sheng et al. (2001) for K-SapCa-Z and Laird et al. (1992) for another Ca-saponite (CEC = 106 cmol/kg) (Figure 3). For further comparison, isothemTS measured by Sawhney and Singh (1997) and Celis et al. (1997) for atrazine sorption by other Ca-saturated reference smectites (CEC = 85 and 76 cmol/kg, respectively) are also included (Figure 3). These six datasets (except that of Celis et al. (1997)) follow the 43 general trend that, for a given exchangeable cation, atrazine sorption tends to increase as layer charge decreases (Laird et al., 1992). (I8 Oti- (14 ~ 0 (12 - v Amount of atrazine sorbed, mg/g clay I I I O 5 10 15 20 01) Equilibrium cone. of atrazine, mg/L K-SapCa-l (this study) K-SapCa-l (Sheng et al., 2001) Ca-SapCa-l (this study) Ca-saponite (Laird et al., 1992) Ca-montmorillonite (Sawhney and Singh, 1997) Ca-montmorillonite (Celis et al., 1997) 04.0.0 Figure 3. Comparison of our sorption results with all earlier studies of atrazine sorption to saponites, as well as two representative montmorillonites. 44 The type of smectite clay does affect triazine pesticide adsorption, as illustrated by the greater atrazine adsorption by saponite than that by beidellite. (Figure 2 a and b). Sorption of atrazine by Cs-SapCa-2 was 9.2 mg/g at the aqueous concentration of 2 mg/L, and is 37 times the sorption by Cs-SBId-l at the same aqueous concentration. Similarly, sorption of atrazine by Na- and Ca-SapCa-2 was greater than the sorption by Na-, and Ca-SBId-l clay, respectively. These observations are again probably due to the availability of sorptive area on clay surfaces. SBId-l clay has a CEC of 129 cmol/kg (Hetzel and Doner, 1993) while SapCa-2 clay has a relatively lower CEC (95 cmol/kg, (Post, 1984)). Generally, clay with a higher CEC retains more hydrated exchangeable cations that cover a larger portion of clay surfaces, hence results in less siloxane surface available for pesticide adsorption, as demonstrated by previous studies of an inverse relationship between clay CBC and pesticide adsorption (Laird et al., 1992; Lee et al., 1990; Sheng et al., 2001; Sheng et al., 2002). In studies of atrazine adsorption by both Ca- and K-saturated smectites, the clay with the highest CEC was consistently the least effective adsorbent for atrazine (Laird et al., 1992; Sheng et al., 2001). Celis et al. (1997) reported the similar observation that Ca-SAz-l (C EC = 120 cmol/kg) sorbed less atrazine than Ca-SWy-l (CEC = 76.4 cmol/kg), which is also consistent with the observation that adsorption of atrazine decreases as clay CEC increases. The current study along with previous studies supports the hypothesis that exchangeable cations (e.g., Ca2+) with larger hydration spheres obscure more of the neutral clay siloxane surface thereby reducing hydrophobic interactions between clay surfaces and pesticides. The lower charge-density clay (SapCa-Z) manifested greater adsorption compared to the higher charge-density SBId-l clay. Lower charged clay apparently has larger exposed areas between 45 exchangeable cations on the neutral siloxane surfaces, which contribute more favorable pesticide adsorptive domains. Crowding interlayer cations upon pesticide adsorption may increase the electrostatic repulsion among these cations, thereby contributing as well to the reduction of pesticide adsorption by the higher charge clay. The site of negative charges on the clay also affects sorption. Permanent negative charge on clay arises from isomorphic substitution in either tetrahedral or octahedral Sites. The CEC of SBId-l is 129 cmol/kg with the negative charges arising from mostly tetrahedral substitution, and SAz-l has a similar CEC (130 cmol/kg) but with the negative charges primarily from octahedral substitution. Data for atrazine sorption by K- SAz-l were reported by Sheng et al. (2001), and the atrazine sorbed by these two KI- saturated clays varied substantially with the much greater sorption for K-SBId-l clay. For example, the amounts of atrazine sorbed by K-SBld-l and K-SAz-l were 0.79 and 0.04 mg/g, respectively, at an aqueous atrazine concentration of 8 mg/L. Since CEC values and surfacie areas of the two clays are similar, differences in sorption result primarily from the sites of negative charges. Compared with the charges from octahedral substitution (SAz-l), the charges from tetrahedral substitution (SBId-l) are localized to fewer surface oxygens (Farmer, 1978; McBride, 1989) resulting in more ordered distribution of exchangeable cations on mineral surfaces (Odom, 1984). These exchangeable cations (at a given hydration state) can always approach the tetrahedrally- substituted sites more closely hence demonstrating a stronger attraction compared with octahedrally-substituted sites (Bleam, 1990). AS a result, SBId-l clay has a smaller clay layer spacing, which favors atrazine adsorption in clay interlayer environments (C happell et al., 2005). In addition, the charge localization of tetrahedral substitution results in a 46 larger portion of neutral siloxane surfaces associated with mineral surfaces, which may provide large neutral surface domains for the adsorption of organic compounds. The K-smectites studied here exhibit interesting sorption behavior. In contrast to the trend discussed above, in which sorption generally decreases as clay layer-charge increases, the higher-charged K-SBId-l sorbs slightly more atrazine than does K-SapCa- 2. Indeed, the sorption by K-SBId-l is equivalent to that of Cs-SBId-l, while sorption by K-SapCa-Z is much below that of Cs-SapCa-Z. One hypothesis is that some C s-SBId-l interlayers collapse and completely dehydrate even in suspension, thus rendering them unavailable for sorption of atrazine. A second hypothesis might be that K-SBId-l sorbs like Cs-SBId-l because the two are Similarly swollen with ~12.5-A dom-spacings: It is well-known (McLean and Watson, 1985) that removal of K+ from an illite clay results in a tetrahedrally substituted clay mineral resembling a high-charged beidellite, while addition of K+ to these weathered illites results in K+ “fixation” to form a mica-like mineral with dehydrated interlayers. The tendency for clay minerals to fix K+ correlates with their layer charge (Sparks and Huang, 1985). While true smectites like K-beidellite should not completely dehydrate and fix K+ (Sparks and Huang, 1985) because their layer charges are lower than those of illites, it is plausible that a hi gher-charged beidellite like K-SBId-l would swell less than a lower-charged smectite like K-SapCa-Z. Thus, despite its high charge, the interlayers of K-SBId-2 should be more likely to retain duo, -spacings near 12.5-A when fully hydrated, and thus be relatively favorable for hydrophobic interactions. Testing of these hypotheses would involve measuring door—spacings of these smectites in aqueous suspension (Chappell et al., 2005), which we hope to do in the future. In the present study we measured the dooj-spacings of air-dried films of K-SBld-2 47 and K-SapCa-2 and found them to be 10.2 and 12.0 A, respectively, so interlayers of the K-beidellite almost completely dehydrate while K-saponite retains one layer of interlamellar water when air-dried. Thus, it is plausible that the beidellite would swell less than the saponite in suspension, as well. The sorption of metribuzin by saponite and beidellite followed the same order as for atrazine, i.e., Cs-saturated > K-saturated z Ca-saturated > Na-saturated clay, but the magnitude of metribuzin sorption was smaller than atrazine sorption by both clays (Figure 2 c and d). This could be due to the higher solubility of metribuzin (1050 mg/L) compared to atrazine (33 mg/L), Since we hypothesize (Boyd et al., 2001) that hydrophobic effects contribute to sorption of neutral organic solutes by smectite clays. Sorption of metribuzin by saponite clay was two to five times (Figure 2c) that by beidellite (Figure 2 d) at a given metribuzin concentration, which may again be due to the comparably lower CEC of saponite as discussed above. The sorption of simazine by these two clays manifested similar results. C s- SapCa-2 adsorbed appreciably higher simazine than K-, Ca- and Na-SapCa-2 (Figure 2 e), and beidellite sorbed considerably less simazine than saponite clay. This again shows that a smectite clay with a lower CEC and/or saturated with a less hydrated exchangeable cation (i.e., CS+) usually demonstrates a greater sorption capacity. Simazine sorption by K-SBId-l behaved like atrazine sorption in that it was almost identical to the sorption by Cs-SBId-l. If hydrophobic effects do contribute to triazine adsorption by smectites, then at a given aqueous triazine concentration, the sorption to a given clay system should have followed the order simazine a: atrazine > metribuzin. Simazine and atrazine are difficult 48 to rank, because atrazine has a slightly higher aqueous solubility (presumably because it is less symmetric) yet also has a larger octanol-water partition coefficient because it contains one more methylene group (Table 1). For the Cs-saponite at 0.1 ppm aqueous triazine, the adsorbed amounts of triazine were 2.0, 2.6, and 0.02 mg/g, for simazine, atrazine, and metribuzin, respectively (Figure 2). For K-saponite at 1.3 ppm aqueous triazine, the adsorbed amounts of triazine were 0.3, 0.1, and 0.03 mg/g, respectively. Similarly, for K-beidellite at 0.5 ppm aqueous triazine, the adsorbed amounts were 0.2, 0.2, and 0.01 mg/ g, respectively. Thus, our limited data are consistent with a hydrophobic component to triazine adsorption by smectites, which implies that one role of the clay interlayer is to partially remove the hydrophobic solute from water (Boyd et al., 2001). Interlayers that perform this role most effectively, such as C s-saponite, should then be the strongest clay mineral sorbent phases. X -ray Diffraction As the atrazine loading increased, the door-spacing (corresponding to the centroid of the diffraction peak) for the air-dried Cs-smectite increased to 12.5 A for the 13.7 g/kg atrazine loading, from 12.1 A for the atrazine-free clay (Figure 4). The increase in basal spacing suggests that atrazine is intercalated in the interlamellar region of the clay, with an increasing fraction of the clay domains at larger d-spacings (~12.5 A) as the atrazine loading increases. This is consistent with our previous observations of intercalated aromatic compounds (Li et al., 2004), in which the dom-spaeings of air-dried K-smectite clay films increase slowly toward ~12.5 A as the aromatic compound loading increases from zero to 200 umol/g clay. Here, the maximum atrazine loading rate is about 65 49 umol/g clay (Figure 4), so sorption is rather small to have a major impact on the d001- spacing (Li et al., 2004), but the systematically increasing dam-spacings are consistent with our hypothesis of interlayer sorption of atrazine. The sorption of metribuzin and simazine did not result in a change in clay basal spacing compared with that in the absence of organics (data not presented). The reason is that the amount of sorbed metribuzin and Simazine in clay interlayer was not sufficient to cause the swelling of clay basal spacing. Li et al. (2004) also reported that at a low amount of organic solute loadings the clay interlayer did not Show the expansion. 12.5 A 12.4 A 12.2 A 12.1 A .1? r 13-7 mg/g 8 . 8 1 .5 8.8 mg/g 0.4 mg/g "“w -- er:- :va Blank 4 6 8 10 12 2 Theta Figure 4. X-ray diffraction patterns of air-dried Cs-SapCa-2 clay films as a function of atrazine sorption. 50 Molecular Simulations The molecular dynamic simulation results offer structural hypotheses for atrazine interactions in the interlayer region of K-SapCa-2. Figure 5 is a representative snapshot taken from a molecular dynamics simulation. As such, it is not a minimum energy structure, but is representative of the dynamic equilibrium. The types of interactions shown in Figure 5 are comparable to others obtained during some 20 similar simulations we have done for atrazine in smectite interlayers. Within a monolayer of interlayer water between the mineral surfaces, atrazine is oriented parallel to the clay basal plane with the aromatic ring interacting with both of the opposing clay siloxane sheets. In this simulated configuration, the negatively charged ring-N atoms of atrazine also form inner- and outer-sphere complexes with multiple K+ ions. The inner-sphere complex shown is a chemically reasonable structure because a K+ ion is able to interact with two (ring N and C 1 substituents) negatively charged (Figure l) and closely spaced sites simultaneously. Our molecular dynamic simulations also support the hypothetical interactions between atrazine and smectite siloxane surfaces that were proposed by Laird (1996): He proposed that the lone pair electrons on the ring N atoms of atrazine interact with water molecules solvating exchangeable cations associated with smectite surfaces; such complexes were indeed observed in our simulations (Figure 5). At the same time, the alkyl-side chains of atrazine molecules were hypothesized (Laird, 1996) to out-compete water molecules for retention on hydrophobic microsites on smectite surfaces. Again, this behavior is observed in the simulations (Figure 5). Such hydrophobic bonding is 51 A2322: 2.523 2: 5:» US$885 59.» u .0 can 6:5" z SEN" O .833 u I 69. n O .59.“ u v— 3 952.8 auto—8 :83 2:. 688188 coma—.85 a... 82.5 3 .598 E .5389. :8: 2::— EE am!— 2: 8 3:22— 295 9.9?— mflo AN -aOnnmtv—v 838:5 1! 2:22:2— ._o .539. coma—L85 2: 5 055...: he noun—35m 383.3 3.322: a .3 “cannaum .m PEME l mcocofioEoEonmoagm 6- Ba 21 wet H 5633 328.885 5388 ecoaamuoafi 52 consistent with the observed decrease in sorption with increasing clay CEC because the amount of neutral siloxane surface on smectites decreases with increasing surface charge density (SCD). On surfaces with low SCD the water molecules at the clay surface are less strongly bound and considered to have a less ordered structure (Gilchrist et al., 1993). Sparingly water-soluble organic compounds (e. g., atrazine, water solubility of 30 mg/L) are more likely to approach and adsorb to surfaces that sorb water less strongly. A lower SCD also means that there are less exchangeable cations competing for adsorption domains in the interlayer region (Sheng et al., 2002). In summary, we observed very strong atrazine sorption by Cs-saponite, and a lesser sorption by K-, Ca-, and Na-saponites. These results are another indication of interplay between the polar functional groups of neutral organic molecules and the hydration of interlayer cations. As the organic functional group becomes more polar, the ability to displace water from clay interlayers becomes stronger, resulting in a substantial adsorption. Thus, pesticides with multiple strongly polar functional groups like —N03 — C=O or -CEN demonstrate a strong affiliation with K- and Cs-smectites (Boyd et al., 2001; Johnston et al., 2002; Sheng et al., 2002) due to pesticide-cation complexation. However, our results indicate that even less polar compounds (e. g., atrazine) can be strongly adsorbed by Cs-saponite, due to the optimal clay interlayer structures, i.e., maintaining ~12.5 A basal spacing and non-overlapping radii of hydration for C s+. Thus, Cs-saponite interlayers contain the largest proportion of relatively hydrophobic sites, in which we hypothesize that triazines can be at least partially dehydrated, and our data are consistent with a hydrophobic component to triazine adsorption by smectites. We also found that the site of negative charge on the clay significantly influenced the 53 retention of pesticide. The smectite clays with the negative charge arising from tetrahedral substitution generally displayed a higher adsorption compared to smectites with octahedral substitution, presumably because tetrahedrally substituted smectites swell less in water and thereby contain narrower slit-pores that favor hydrophobic sorption. However, if the clay layer charge is large, adsorption of triazine herbicides by even tetrahedrally substituted smectites can be sharply reduced, because hydration radii for the greater number of exchangeable cations begin to overlap. Apparently, an optimal inorganic sorbent for triazine herbicides should be a Cs+-saturated smectite with a low layer charge resulting from tetrahedral substitution. These criteria maximize adsorption domains parallel to the clay surfaces while optimizing (near 12.5 A) the adsorption domains perpendicular to the clay surfaces. If true, these hypotheses should apply broadly to hydrophobic organic solutes. 54 LITERATURE CITED Albanis, T.A., P.J. Pomonis, and AT. Sdoukos. 1988. Movement of methyl parathion, lindane, and atrazine through lysimeters in field conditions. Toxicological and environmental chemistry 17:35-45. Armstrong, DE, and G. Chesters. 1968. Adsorption catalyzed chemical hydrolysis of atrazine. Environmental Science & Technology 22683-689. Bleam, W.F. 1990. The nature of cation-substitution sites in phyllosilicates. Clays and Clay Minerals 38:527-536. Boyd, S.A., G.Y. Sheng, B.J. Teppen, and OJ. Johnston. 2001. Mechanisms for the adsorption of substituted nitrobenzenes by smectite clays. Environmental Science & Technology 3524227-4234. Celis, R., J. Comejo, M.C. Hermosin, and W.C. Koskinen. 1997. Sorption-desorption of atrazine and Simazine by model soil colloidal components. Soil Science Society of America Journal 612436-443. Chappell, M.A., D.A. Laird, M.L. Thompson, H. Li, B.J. Teppen, V. Aggarwal, C.T. Johnston, and SA. Boyd. 2005. Influence of smectite hydration and swelling on atrazine sorption behavior. Environmental Science & Technology 39:3150-3156. Chiou, C.T., L.J. Peters, and V.H. Freed. 1979. Physical Concept of Soil-Water Equilibria For Non-Ionic Organic-Compounds. Science 206:831-832. Chiou, C .T., D.E. Kile, D.W. Rutherford, G.Y. Sheng, and SA. Boyd. 2000. Sorption of selected organic compounds from water to a peat soil and its humic-acid and humin fractions: Potential sources of the sorption nonlinearity. Environmental Science & Technology 3421254-1258. Comber, S.D.W. 1999. Abiotic persistence of atrazine and simazine in water. Pesticide Science 55:696-702. Evangelou, VP. 1998. Environmental soil and water chemistry 2 principles and applications John Wiley & Sons, New York. Farmer, V.C. 1978. Water on particle surfaces, p. 405-448, In Greenland and Hayes. eds. The chemistry of soil constituents. John Wiley and Sons, New York, NY. Gamble, D.S., and S.U. Khan. 1985. Atrazine hydrolysis in soils: Catalysis by the acidic functional groups of fulvic acid. Canadian Journal of Soil Science 65:435-443. 55 Gilchrist, G.F.R., D.S. Gamble, H. Kodama, and S.U. Khan. 1993. Atrazine interactions with clay-minerals — kinetics and equilibria of sorption. Journal of Agricultural and Food Chemistry 41:1748-1755. Grundl, T., and G. Small. 1993. Mineral contributions to atrazine and alachlor sorption in soil mixtures of variable organic carbon and clay content. Journal of Contaminant Hydrology 142117-128. Hetzel, F., and HE. Doner. 1993. Some colloidal properties of Beidellite — Comparison with low and high charge montmorillonites. Clays and Clay Minerals 41:453-460. Huang, P.M., R. Grover, and RB. McKercher. 1984. Components and particle-size fractions involved in atrazine adsorption by soils. Soil Science 138220-24. Huang, W.L., T.M. Young, M.A. Schlautman, H. Yu, and W.J. Weber. 1997. A distributed reactivity model for sorption by soils and sediments .9. General isotherm nonlinearity and applicability of the dual reactive domain model. Environmental Science & Technology 31 :1703-1710. Johnston, C.T., G. Sheng, B.J. Teppen, S.A. Boyd, and M.F. De Oliveira. 2002. Spectroscopic study of dinitrophenol herbicide sorption on smectite. Environmental Science & Technology 3625067-5074. Karickhoff, SW. 1984. Organic pollutant sorption in aquatic systems. Journal of Hydraulic Engineering 110:707-735. Kleineidam, S., H. Rugner, B. Ligouis, and P. Grathwohl. 1999. Organic matter facies and equilibrium sorption of phenanthrene. Environmental Science & Technology 33:1637-1644. Koskinen, W.C., and SS. Harper. 1990. The retention process: Mechanisms, p. 51-78, In H. H. Cheng, ed. Pesticides in the soil environment: Processes, impacts, and modeling. Soil Science Society of America, Madison, WI. Laird, DA. 1996. Interactions between atrazine and smectite surfaces, p. 86-100, In M. T. Meyer and E. M. Thurman, eds. Herbicide Metabolites in Surface Water and Groundwater, Vol. 630. American Chemical Society, Washington, DC. Laird, D.A., E. Barriuso, R.H. Dowdy, and W.C. Koskinen. 1992. Adsorption of atrazine on smectites. Soil Science Society of America Journal 56262-67. Lee, J .-F ., M.M. Mortland, C.T. Chiou, D.E. Kile, and SA. Boyd. 1990. Adsorption of benzene, toluene, and xylene by two tetrarnethylarnmonium-smectites having different charge densities. Clays and Clay Minerals 38:113-120. Li, H., B.J. Teppen, C.T. Johnston, and SA. Boyd. 2004. Thermodynamics of nitroaromatic compound adsorption from water by smectite clay. Environmental Science & Technology 38:5433-5442. 56 Li, H., G.Y. Sheng, B.J. Teppen, C.T. Johnston, and SA. Boyd. 2003. Sorption and desorption of pesticides by clay minerals and humic acid-clay complexes. Soil Science Society of America Journal 67:122-131. Maple, J .R., M.-J. Hwang, T.P. Stockfish, U. Dinur, M. Waldman, C.S. Ewig, and AT. Hagler. 1994. Derivation of Class 11 force fields. 1. Methodology and quantum force field for the alkyl functional group and alkane molecules. Journal of Computational Chemistry 15:162-182. McBride, MB. 1989. Surface chemistry of soil minerals, p. 35-88, In J. B. Dixon and S. B. Weed, eds. Minerals in soil environments, 2nd ed. Soil Science Society of America, Madison, WI. McBride, MB. 1994. Environmental chemistry of soils Oxford University Press, New York. McLean, E.O., and ME. Watson. 1985. Soil measurements of plant-available potassium, p. 277-305, In R. D. Munson, ed. Potassium in Agriculture. American Society of Agronomy, Madison, WI. Means, J .C., S.J. Wood, J .J . Hassett, and W.L. Banwart. 1982. Sorption of amino- and carboxy-substituted polynuclear aromatic hydrocarbons by sediments and soils. Environmental Science & Technology 16293-98. Accelrys, Inc. 2000. Cerius2 Molecular Simulation Software, Release 4.2. Accelrys, Inc., San Diego, CA. Noblet, J .A., LA. Smith, and 1H. Suffet. 1996. Influence of natural dissolved organic matter, temperature, and mixing on the abiotic hydrolysis of triazine and organophosphate pesticides. Journal of Agricultural and Food Chemistry 4423685- 3693. Odom, I.E. 1984. Smectite clay minerals: properties and uses. Philos Trans Royal Soc London Ser A3112391-409. Post, J .L. 1984. Saponite from near Ballarat, California. Clays and Clay Minerals 32:147- 153. Sannino, F ., M.T. Filazzola, A. Violante, and L. Gianfreda. 1999. Adsorption-desorption of simazine on montmorillonite coated by hydroxy aluminum species. Environmental Science & Technology 3324221-4225. Sato, T., T. Watanabe, and R. Otsuka. 1992. Effects of layer charge, charge location, and energy change on expansion properties of dioctahedral smectites. Clays and Clay Minerals 40:103-113. Sawhney, B.L., and SS. Singh. 1997. Sorption of atrazine by Al- and Ca-saturated smectite. Clays and Clay Minerals 452333-338. 57 Sheng, G.Y., C.T. Johnston, B.J. Teppen, and SA. Boyd. 2001. Potential contributions of smectite clays and organic matter to pesticide retention in soils. Journal of Agricultural and Food Chemistry 49:2899-2907. Sheng, G.Y., C.T. Johnston, B.J. Teppen, and SA. Boyd. 2002. Adsorption of dinitrophenol herbicides from water by montmorillonites. Clays and Clay Minerals 50225-34. Sonon, LS, and AP. Schwab. 1995. Adsorption characteristics of atrazine and alachlor in Kansas soils. Weed Science 43:461-466. Sparks, D.L., and RM. Huang. 1985. Physical chemistry of soil potassium, p. 201-276, In R. D. Munson, ed. Potassium in Agriculture. American Society of Agronomy, Madison, WI. Sun, H. 1998. COMPASS: An ab initio force-field optimized for condensed-phase applications - Overview with details on alkane and benzene compounds. Journal of Physical Chemistry B 102:7338-7364. Suquet, H., and H. Pezerat. 1987. Parameters influencing layer stacking in saponite and vermiculite: A review. Clays and Clay Minerals 35:353-362. Talbert, R.E., and OH. Fletchall. 1965. The adsorption of some s-triazines in soils. Weeds 13:46-52. Teppen, B.J., K. Rasmussen, P.M. Bertsch, D.M. Miller, and L. Schafer. 1997. Molecular dynamics modeling of clay minerals .1. Gibbsite, kaolinite, pyrophyllite, and beidellite. Journal of Physical Chemistry B 101 : 1579-1 587. Weber, W.J., and W.L. Huang. 1996. A distributed reactivity model for sorption by soils and sediments .4. Intraparticle heterogeneity and phase- distribution relationships under nonequilibrium conditions. Environmental Science & Technology 302881 - 888. Xia, GS, and WP. Ball. 2000. Polanyi-based models for the competitive sorption of low- polarity organic contaminants on a natural sorbent. Environmental Science & Technology 34:1246-1253. 58 CHAPTER III ATRAZINE INTERACTIONS WITH SMECTITE CLAYS: MOLECULAR SIMULATION APPROACH 59 Atrazine interactions with smectite clays: Molecular Simulation approach Introduction Atrazine is one of the most commonly used agricultural herbicides in the world and has been applied for more than 40 years by com and sorghum growers in the United States (Ribaudo and Bouzaher, 1994). In 1994, over 47 million pounds of atrazine were applied nation-wide covering about 77 % of the area cropped under com-soybean rotation (Pike and Knake, 1996). Atrazine is also one of the most common herbicide residues found in groundwaters of Midwest (Blanchard and Donald, 1997; Burkart et al., 1999; Kolpin et al., 1998), and is detected 10 to 20 times more frequently than the next most commonly detected pesticide (Belluck et al., 1991). The presence of atrazine in groundwater is a concern because atrazine is a potential carcinogen (Biradar and Rayburn, 1995), and a suspected endocrine disrupter. Toxicological concerns have resulted in the EPA setting a maximum amount of 3 ug/L of atrazine in drinking water in the US. Despite restrictions, atrazine exceeds the maximum contaminant level of 3ppb at certain times in several states in the US. It has been shown that adsorption of organics on smectite clays is an important phenomena that controls the bioavailability, persistence and movement of organics in soils and subsoils (Boyd et al., 2001; Laird et al., 1992; Li et al., 2003; Sheng et al., 2001). The sorption capacity of smectites for organics varies widely depending on the surface charge density of the smectite (Laird et al., 1992), the nature of the adsorbed 60 cation (Sheng et al., 2001), the pH of the soil (Vaz et al., 1997), and the location of negative charge on the smectite. In parallel with atrazine usage in agriculture, researchers have studied atrazine adsorption by soils and soil components for more than 40 years. Due to its long-term agronomic importance and its leaching risk, atrazine sorption processes in soils have probably been studied more than those of any other organic solute. Most studies, of course, have been phenomenological in nature, striving to characterize the extent of sorption and correlate sorption with known properties of atrazine and its related triazines, with soil properties, and with solution conditions such as pH. However, there has been continual and intense interest in mechanistically understanding the forces that drive triazine sorption. Already by 1962, Frissel and Bolt (1962) described triazine sorption to montmorillonite as a cation exchange reaction within a few pH units of each triazine pKa. Furthermore, they speculated that >90% of triazine molecules would be adsorbed by clay-rich soils, even at neutral pH, through the action of van der Waals forces between triazines and clay siloxane surfaces (Frissel and Bolt, 1962). Weber et al. (1965) used X- ray diffraction to Show that prometone (a methoxy— rather than chloro-triazine, with pKa of 4.3) intercalated between clay layers. Sorption of 150 to 300 umol/g caused the dry clay to swell to ~13 A, consistent with triazine rings oriented parallel to the Silicate sheet. Sorption of 450 umol/g caused clay swelling to ~18 A, and the authors used molecular models to postulate triazine molecules either perpendicular to the clay surfaces or arranged in double stacks in the interlayer. Most sorption to the montmorillonite was by cation exchange, since the experiments were done just 1.7 pH units above the pKa, 61 though the authors postulated neutral species sorption via H-bonding of triazines to siloxane surface 0 atoms of the clay (Weber et al., 1965). Bailey et al. (1968) studied sorption of six triazines by Na-montmorillonite and found that sorption was generally larger for the more soluble triazines, the opposite of the expected behavior if a nonspecific, hydrophobic adsorption mechanism were operant (Hance, 1967). However, triazine cation formation can only be discounted for the three low-pKa triazines studied (including atrazine), and for these compounds there was not a consistent relationship between sorption and aqueous solubility (Bailey et al., 1968). For neutral triazines interacting with clay minerals, they hypothesized that possible adsorption mechanisms might include van der Waals interactions and H-bonding between amino N-H groups and clay siloxane O atoms. For many years, though, the dominant mechanism for atrazine adsorption by clay minerals was thought to be cation exchange, because no sorption of uncharged (“molecular”) atrazine by clays was conclusively demonstrated. However, Laird et al. (1992) argued that, at neutral pH, molecular atrazine was the species sorbing to a variety of smectites, and they showed that adsorption was inversely correlated to smectite layer charge. In a later study, Barriuso et al. (1994) measured desorption from many of these same smectites, showed that there was little hysteresis at work, and concluded that adsorption of molecular atrazine to smectites is generally reversible. As a result, they hypothesized that the atrazine sorption mechanisms were mainly van der Waals or hydrogen bonding to smectite surfaces. The finding that atrazine adsorption was inversely correlated with smectite layer charge (Laird et al., 1992) implied a hydrophobic sorption mechanism, as had been 62 asserted by Jaynes and Boyd (1991) to explain their observations that benzene adsorption by smectites increased with decreasing clay layer charge. Expandable clay minerals are known to be strongly hydrophilic on a macroscopic scale (Mooney et al., 1952; Xu et al., 2000). However, on a molecular scale there are uncharged regions between the charged sites on smectite siloxane surfaces, so smectites with lower layer charges will have a greater proportion of uncharged surface. J aynes and Boyd (1991) showed that, when smectites are saturated by organic rather than inorganic cations, the uncharged regions of the smectite surface will preferentially adsorb benzene versus water from aqueous solution, so they described these uncharged regions as hydrophobic. In smectites, these hydrophobic regions might comprise more than 40% of the total surface area (Laird, 1996). In other contexts, the sorption of nonpolar, nonionic organic solutes is known to be controlled by “hydrophobic effects,” a combination of relatively small van der Waals bonding forces and a substantial entropic gradient that drives the organic molecules out of aqueous solution (Israelachvili, 1992; Tanford, 1980; Voice and Weber, 1983). Previous studies (Boyd et al., 2001; Curtiss et al., 1986; Laird and Fleming, 1999; Mader et al., 1997; Piatt et al., 1996; Rebhun et al., 1992; Schwarzenbach and Westall, 1981; Van Blade] and Moreale, 1974) have observed that the association of organic solutes with an inorganic surface increased with decreasing aqueous solubility of the compounds. Laird (1996) described the reversible sorption of molecular atrazine as a hydrophobic process, in which the hydrophobic alkyl side-chains of atrazine interact with the hydrophobic microsites located between smectite charge sites. At the same time, he 63 hypothesized (Laird, 1996), N atoms of the triazine ring may accept H-bonds from water molecules that solvate nearby cations, further stabilizing the atrazine-clay complex. Laird (1996) predicted that, if the hydrophobic sorption mechanism for atrazine on smectites was true, then triazine sorption on montmorillonite should increase with the length of the alkyl side-chains, which had been seen in earlier data (Bailey et al., 1968; Weber, 1970) but erroneously ascribed to pKa differences. Laird and Fleming (1999) tested this hypothesis by comparing the sorption of molecular pyridine and butylpyridine at high pH: The sorption of butylpyridine was far larger than that of pyridine, thus supporting the concept of hydrophobic contributions to organic adsorption on smectites. Spectroscopic characterization of atrazine sorption mechanisms has been limited to organic matter as a sorbent phase (Chien and Bleam, 1997; Welhouse and Bleam, 1992; Welhouse and Bleam, 1993), and we are aware of no spectroscopic studies of molecular atrazine in clays. Protonation (Cruz et al., 1968) and hydrolysis (Russell et al., 1968) of triazines on smectites have been studied by FTIR spectroscopy, but not the bonding environment of the neutral triazine. Since atrazine adsorbs to clays as the molecular species at neutral pH (Laird et al., 1992), we wish to understand its mechanism of bonding. It is difficult to visualize the possible interactions between clay surfaces, water, and organics at the molecular level. Molecular dynamics simulations can help in understanding these interactions by providing an objective, energy-based approach to modeling the structural arrangements likely to occur and persist in smectite interlayers. Ideally, simulation results would be coupled with other molecular-scale data such as vibrational, NMR, neutron scattering, or EXAF S spectroscopies (Teppen et al., 1997) in order to employ multiple perspectives. 64 In the present case, we only have bulk adsorption data plus X-ray diffraction data to constrain the smectite interlayer spacing. The objective of the present study, therefore, was to couple molecular dynamics simulations with the available data in order to explore the behavior of atrazine on hydrated smectite clay surfaces. Materials and Methods Molecular dynamics simulations of atrazine-clay complexes in the interlayer region were performed to explore the interactions between atrazine, exchangeable cations, and water. In this approach, electrons are not treated explicitly, but their effects are parameterized into classical (that is, non-quantum) mechanical energy functions comprising a force field. Models for three different smectites were created, representing a Wyoming montmorillonite (lower charge due to octahedral substitution), an Arizona montmorillonite (higher charge due to octahedral substitution), and a California saponite (moderate charge due to tetrahedral substitution). The three correspond to the reference clay minerals SWy-2, SAz-l, and SapCa-2, respectively, for which we have atrazine sorption data. Homoionic versions of each smectite with K+, CS+, or Ca2+ in the interlayer were simulated. The composition of the model K-saturated Wyoming smectite (K—SWy) was K7(A142Mg6)(Si95A11)Ozao(OH)48 , while that of Cs-smectite (Cs-SWy) was Cs7(A142Mg6)(S195A11)0240(OH)43. The CEC of our model K-SWy was 79 cmol kg", and the comparable experimental value is 83 cmol kg'l (Sposito et al., 1983). We also created models for K- and Ca-saturated Arizona smectite (SAz-l). The composition of this model K-smectite (K-SAz) was K12(Al3(,Mg12)(Si%)0240(OH)48 . while that of C a- 65 smectite (Ca-SAz) was Ca6(A136Mg12)(Si96)Ozao(OH)43. The K-SAz was a high-charged smectite, with a modeled CEC of 132 cmol kg’l, which compares with the experimental value of 130 cmol kg'1 (Laird et al., 1992). The model for K-saponite had composition K9(S137A19)Mg720240(OH)43 with a CEC of 95 cmol/kg, in agreement with its experimental CEC (Post, 1984). The smectite surfaces were constructed using methods previously detailed (Teppen et al., 1997), in which the muscovite mica structure (Brigatti and Guggenheim, 2002) was the starting point and Al-for-Si or Mg-for-Al substitutions were made to charge the layers. One molecule of atrazine was added to each unit cell of the smectite, with the atrazine ring perpendicular to the basal plane of the clay. The simulated loading rate was therefore ~24 mg atrazine per gram clay for each atrazine-smectite system. This is larger than our experimental loading rates, which are all less than 14 mg/ g (see Chapter 2), but lower simulated loading rates would have required larger unit cells and therefore much slower simulations. As will be shown below, the present unit cell is large enough to allow simulation of relatively isolated atrazine molecules, while small enough to allow many different systems to be sampled. Water molecules were then added (in random interlayer positions) so that constant-pressure simulations resulted in 12.5, 15.5 or 18.5 A clay layer spacings. One example of an initial configuration is shown in Figure 1. 66 Figure l. A typical starting configuration before beginning the molecular dynamics simulation. Here, atrazine has been placed into the interlayer of Cs-SWy-2 with 55 water molecules. The equilibrium dam-spacing will eventually be 12.3 A. The energies and forces were computed using a force field developed for clays (Teppen et al., 1997) combined with the polymer consistent force field or pcff (Maple et al., 1994; M81, 2000) for the organics and water. As a further modification, charges on the atrazine were assigned using the COMPASS force field (Sun, 1998), which optimized atrazine atomic charges for the condensed phase systems. In previous work (Boyd et al., 2001; Sheng et al., 2002) we have found COMPASS charges to more faithfully reproduce quantum mechanical charges for nitrogenous organic molecules than do PCFF charges. In aqueous interfacial systems, the charges on the atoms of organic solutes play a critical role in determining organic solute interactions with the clay and interlayer water and cations. In classical molecular dynamics, the nuclear charge and the local electron density 67 are subsumed into one partial atomic charge and these charges for atrazine are illustrated inFigure2. Tl C1 N1/ \\N1 H ’I’r’ / ““ H arge ‘\N2/ N1 \ / 1 Cl +0.583O H N2 C2 +0.5637 2 C +01048 a \ . . 2\ 3/ /H7- Ca 01590 H2\c,—--——C4—H2 C5 +0.1578 \ N. -O.4810 c4 / 112 N2 -05422 H2// \H2 H1 +0.24g7 H2 C4—H2 H2 +0.0530 H2 \H C1 01020 2 Figure 2. Partial atomic charges on the atoms of the atrazine molecule, as used in all simulation studies. The total charge on the molecule sums to zero. Commercial software (M81, 2000) was used to compute the classical molecular dynamics, with a timestep of 0.2 fs. Ewald summation was used for both electrostatic and van der Waals forces (Karasawa and Goddard, 1989). No constraints were imposed on our systems during their molecular dynamics except for the contents of the simulation cell: All atoms were completely free to move, and unit cell shape and volume were unconstrained. The molecular dynamics simulations were run in the NPT (constant number of atoms, pressure, and temperature) ensemble for at least 50 ps, much longer 68 than the time required for the system volume and energy to equilibrate. The unit cell volume equilibrates primarily through changes in the dew-spacing, which is directly comparable to X-ray diffraction data. Note that the unit cell has a constant composition. We used periodic boundary conditions, which allow our small unit cell to more faithfully represent a condensed phase. As such, atoms are free to diffuse out of the central unit cell, but a copy of the atom will simultaneously diffuse in from the opposite side, maintaining the constant composition. Results and Discussion Within approximately 5 ps of simulation time, the volume of the atrazine-clay- water systems was typically equilibrated, while the energy was usually equilibrated within the first 10 ps. Thus, the final 40 ps of each Simulation were used as a sampling of the equilibrium dynamics of the model system. In many cases, systems of the same composition (or varying by one to a few water molecules) were simulated multiple times and the equilibrium structures were quite consistent. The molecular dynamic simulation results offer structural hypotheses for atrazine interactions in the interlayer regions of three smectites, saturated with three different cations, at three different water contents. Representative snapshots of the various equilibrated atrazine-water-clay systems taken from molecular dynamics simulations are shown in Figures 3-10. These snapshots are not minimum energy structures, but are representative of the dynamic equilibrium. Atrazine is thought to interact with smectite surfaces through a variety of bonding mechanisms (Laird, 1996; Laird and Sawhney, 2002). Our molecular simulations revealed possible orientations of atrazine in the interlayers of our simulated smectites 69 with one, two, and three layers of interlayer water molecules. Figure 3 allows for comparison of the effect of water content on the atrazine molecular environment in K- SWy-2. The swelling of the smectite is a critical factor controlling atrazine adsorption: Chappell et al. (2005) recently showed that a hydrated K-smectite consisting of interstratified monolayers and bilayers (i.e., Figure 3A and 3B) adsorbed more than ten times the atrazine as did the same K-smectite when swelled to a trilayer (i.e., Figure 3C). The authors (Chappell et al., 2005) used drying-induced hysteresis to cause swelling differences, and were able to measure the K-smectite dam-spacings in the suspension phase, so these spacings represent the interlayer hydration at which atrazine sorption actually occurs. These recent measurements (Chappell et al., 2005) of the effect of clay swelling on atrazine adsorption help to explain previous discoveries that addition of NHa+ fertilizer salts to soils increases atrazine adsorption (Clay et al., 1998) and that decreasing the soil water content leads to increased atrazine sorption (Dao and Lavy, 1978). Structural trends for the monolayer, bilayer, and trilayer cases and their relation to differences in sorption will be discussed in more detail below. 70 4.» Ii. q. V. a- 1, I! 1’&' :L'i' t'ir' I'a'é' 'ir 71 Figure 3. Snapshot showing the orientation of atrazine within the interlayer of a K- saturated smectite like SWy-2 containing (A) one, (B) two, and (C) three layers of hydration water (red = oxygen, white = hydrogen, blue = nitrogen, gray = carbon, green = interlayer K+ cations). Images on the left depict a side-view (x,z plane) of the atrazine-smectite interlayer complex while the images on the right depict the same from a top-view (x,y plane), having removed the silicate layers. The simulated dam-spacings for A, B, and C are 12.4, 15.5, and 17.3 A, respectively. Figure 4 illustrates atrazine structural relationships to cations and water within the interlayer regions of four different simulated smectites, all with door-spacings near 12.5 A. In such smectites containing a monolayer of interlayer water, atrazine is necessarily oriented parallel to the basal surfaces and interacts directly (no intervening water molecules) and simultaneously with both opposing basal surfaces. In this simulated configuration, with its narrow door-spacing and limited volume to accommodate atrazine and several cations, the negatively charged ring-N and -C1 atoms of atrazine could hardly escape contact with multiple K+ (or CS+) ions. Figure 4. Snapshots of molecular dynamic simulations of atrazine in the interlayer regions of four different homoionic smectites, (A) K-SWy-Z, (B) Cs-SWy-Z, (C) K- SAz-l, and (D) K-SapCa-Z. Clay layers were parallel to the page and have been removed in order to expose the interlayer complexes. The atom coloring scheme is K = green, 0 = red, H = white, C =gray, N =b1ue, and CI = green (associated with the atrazine molecule). 73 In the snapshots of Figure 4, which are typical of their respective simulations, the negatively charged ring-N and -Cl atoms of atrazine thus tended to form both inner- and outer-sphere complexes with multiple interlayer cations. The —C1 substituent of the atrazine ring often coordinated directly to a single K+ (or Cs+) ion (Figure 4) with average K (or Cs-) -Cl distances of 3.2 A :t 0.3 A (for K- SAz), 3.9 d: 0.2 A (for Cs-SWy), 4.2 :1: 0.2 A (for K-SWy), and 4.0 :h 0.1 A (for K- SapCa-Z). For example, in the Cs-SWy—2 system, approximately 5 ps of simulation time passed before Cs+ and -Cl approached each other to form their inner sphere complex. Once formed, this complex remained stable over the subsequent 45 ps of simulated time (Figure 5). o I 7 I I I 0 1 0 20 30 40 50 Figure 5. Trajectory of the Cs — Cl inner sphere nonbonded complex in the simulated atrazine-Cs-SWy-Z system represented by Figure 4B. 74 Secondly, it was also typical that each ring -N- atom adjacent to the —Cl substituent of atrazine coordinated directly (inner-sphere) to a single KI (or Cs+) ion (Figure 4), with an average K (or Cs-) -N distance of 2.7 A d: 0.2 A (for K-SAz), 3.1 i 0.2 A (for Cs-SWy), 3.7 a 0.2 A (for K-SWy), and 3.2 :1: 0.2 A (for K-SapCa-2). The ring —N- atom in the 5-position seems to be hindered from engaging in such complexes by the alkyl substituents in the 4- and 6-positions. Our molecular dynamic simulations also support the hypothetical interactions between atrazine and smectite siloxane surfaces that were proposed by Laird (1996): He proposed that the lone pair electrons on the ring N atoms of atrazine interact with water molecules solvating exchangeable interlayer cations; such complexes were indeed observed in our simulations. Transient states existed in which ring —N atoms (in the 1- and 3-positions) were indirectly coordinated to K+ (or CS+) ions through the intermediation of water. For example, the outer-sphere C s -- N complex in our model SWy-2 (Figure 4B) exhibited average distance of 5.4 i 0.3 A. Figure 4 shows that, for three of the four Cs- and K—smectites that we simulated, it was typical for two atrazine functional group sites (ring-N and -C1) to simultaneously interact with one interlayer cation. These inner-sphere complexes seem to be chemically reasonable structures because a cation is able to interact with two negatively charged (- 0.48 and -0.10 from Figure 2) and closely spaced sites simultaneously. We were somewhat surprised to see such complexes form, because 0 atoms of water have charges of -0.82 in our system and therefore ought to out-compete atrazine for coordination sites around cations. However, we see cations “complexed” to -Cl, -N- atoms, and even methyl groups of atrazine (Figure 4). 75 The steric conditions of crowded interlayers at 12.5-A door-spacing may play a role in formation of such “complexes.” For example, we have seen that the K -- Cl complexes in SAz-l and SWy-2 had lengths of 3.2 A i 0.3 A and 4.2 :t 0.2 A, respectively. Also, K -- N distances were 2.7 A d: 0.2 A and 3.7 i 0.2 A, respectively. Part of the force driving the K-atrazine complexes closer in SAz-l may be mutual repulsions of all the cations, since it is a higher-charged interlayer environment than is SWy-2. It is visually apparent (Figure 4C) that atrazine has disrupted the average spacing of the 12 K+ ions per unit cell in SAz-l, so another reason for shorter K-atrazine complexes may simply be that atrazine is interfering with optimal interactions between interlayer cations and charged sites in the clay layers. In the simulations, we force atrazine into the K-SAz-l interlayer, but in reality it does not need to adsorb. Thus (Sheng et al., 2001), K-SAz-l (Figure 4C) adsorbs only about 1/6"1 the atrazine as does K-SWy-2 (Figure 4A), presumably because the former has higher charge, so has a more hydrophilic interlayer, so atrazine has a harder time competing with water for interlayer sites (Laird, 1996). Indeed, Laird et al. (1992) showed that atarazine sorption by Ca- smectites was inversely correlated with smectite layer charge. Figures 4A and B compare the environment of atrazine in K- versus Cs-SWy-2 montmorillonite. Note that K+ cations near atrazine (Figure 4A) are more likely to have water molecules competing for them, disrupting the K-atrazine complexes, while C s+ cations (Figure 4B) are able to form two inner-sphere complexes with atrazine. This is consistent with the less favorable hydration energy of Cs+. As a result, the Cs-smectite interlayer is less hydrophilic and atrazine competes better for interlayer sites. For example, using typical estimates (Evangelou, 1998) for the hydrated radii of C 8+ (3.6 A) 76 and K+ (5.3 A), along with an assumed 700 mz/g basal surface area and monolayer hydration status, we compute that only 56% of the Cs-SWy-2 interlayer surface is occupied by hydrated cations while 122% of K-SWy-2 is occupied. Thus, more of the relatively hydrophobic surface should be available for atrazine sorption in the C s-system. We know (Chapter 2) that Cs-SapCa-2 sorbs about 50 times as much atrazine as K- SapCa-Z, and the saponite’s layer charge is close to that of SWy-2, but we do not yet have the sorption data for Cs-SWy-2. Given Laird’s (Laird et al., 1992 and Laird, 1996) hypothesis that lower-charged smectites adsorb more atrazine because the surface charge sites are more widely separated, so a larger fraction of the smectite surface is hydrophobic, we looked to see if indeed atrazine sorbs to hydrophobic nanosites in the smectite interlayers. Figure 6 shows that most of the atrazine molecule occupies a more-hydrophobic surface site between charge sites in K-SWy-2. However, given the rather regular charge spacing in our model SWy-2, atrazine cannot quite fit completely between charged sites. For the Cs-SWy-2 (Figure 6), the atrazine ring resides almost directly over a negative site but the two alkyl sidechains occupy more hydrophobic sites. 77 Figure 6. Representative snapshot of the relationships of atrazine and cations to charged sites in model SWy-2 with a monolayer of water for the K— (left) and Cs- saturated (right) cases. Blue octahedra are the locations of Mg-for-Al octahedral substitution sites and pink tetrahedral are the locations of Al-for-Si tetrahedral substitution sites in the montmorillonite layer. The upper clay layer and the rest of the lower clay layer have been removed for clarity. For the more expanded smectites with two or three layers of interlayer water molecules, fewer atrazine-cation complexes formed, and the atrazine typically inclined at an acute angle relative to the basal surfaces (Figure 3B and 3C). The two alkyl side chains tended to interact directly with one basal surface each. The rest of the atrazine molecule was forced to interact with interlayer water molecules. Figure 7 shows snapshots from four simulations of atrazine in the interlayers of hydrated smectites (both SWy and 8A2), each containing a bilayer of interlayer water between the mineral surfaces. The equilibrium door—spacings of these systems were 15.5 i 0.1A, which is a realistic interlayer spacing for smectites having two layers of 78 A :IW'IJEH *2 {“2 u "tii QIp-j‘"~1 4. ti E: ‘ VI Figure 7. Snapshots of molecular dynamic simulations of atrazine in the interlayer regions of four different homoionic smectites, (A) K-SAz-l, (B) Ca-SAz-l, (C) K- SWy-2, and (D) Cs-SWy-Z. Clay layers were parallel to the page and have been removed in order to expose the interlayer complexes. The atom coloring scheme is K = green, 0 = red, H = white, C =gray, N =b1ue, and CI = green (associated with the atrazine molecule). 79 interlayer water. In this environment atrazine did not orient its triazine ring parallel to the clay surface during 50 ps of simulation time. In the Ca-saturated SAz-l (Figure 7B), the negatively charged functional groups of atrazine formed only outer-sphere complexes with interlayer Ca2+ ions. In the K-SWy-2 bilayer system, the negatively charged ring-N atoms of atrazine were again able to form both inner- and outer-sphere complexes with K+ ions. For example, the distance between one of the interlayer K and ring-N atoms of atrazine remained stable around 3.1 i 0.3A over almost all of the simulation time (Figure 8). For the CS-SWy-Z bilayer system, just one interlayer Cs+ formed an inner-sphere complex with a ring-N atom of the atrazine, compared with two inner-sphere and one outer-sphere complex in the monolayer Cs-hydrate (Figure 4B). 0 1 0 20 30 40 50 Figure 8. Trajectory of the K - N inner sphere nonbonded complex in the simulated atrazine-K-SWy-2 system represented by Figure 3B and 7C. 80 Again, it is of interest to observe the spatial relations between adsorbed atrazine, interlayer cations, and charge substitution sites in the clay layers. Figure 9 shows that, in the K-SWy-2 bilayer system, one alkyl chain of atrazine lies in one water layer (on one basal surface) and the other alkyl chain lies in the second layer of interlayer water (on the opposite basal surface). For the Cs-SWy bilayer system, too, the two alkyl side chains interact directly (no intervening water molecules) with one basal surface each (Figure 9). ‘I I:“"" C‘Kf“ ‘5. Sta F A 1r it A 1r 1t .; ‘83.wa I gist-7 71:84 1:50;.(3 #313011 «A 'afit 1,... \ Figure 9. Representative snapshots of the relationships of atrazine and cations to charged sites in model SWy-2 with a bilayer of water for the K- (left) and Cs- saturated (right) cases. Blue octahedra are the locations of Mg-for-AI octahedral substitution sites and pink tetrahedral are the locations of AI-for-Si tetrahedral substitution sites in the montmorillonite layer. The lower clay layer and the rest of the upper clay layer have been removed for clarity. Inspection of the molecular simulation results indicates that the two alkyl chains of atrazine tend to interact with uncharged portions of the montmorillonite basal surface, just as hypothesized by Laird (1996) and Laird and Fleming (1999). Perhaps if the smectite charge distribution were more heterogeneous so that some larger uncharged 81 domains were created, atrazine would be stable when adsorbed flat on one surface. This could be a subject for future simulations. Finally, further simulations of atrazine in the interlayer space of SWy-2 or SAz-l modeled even wetter conditions. Here, a trilayer of interlayer water formed between the mineral surfaces. The equilibrium door-spacing was 18.5 i 0.1A for Ca-SAz-l and 17.3 :t 0.1A for K—SWy-2. These are also realistic interlayer spacings for smectites having three layers of interlayer water. In this environment atrazine is again unable to orient parallel to the clay surface during the 50-ps simulations. Rather, the situation is similar to the bilayer case, with atrazine inclined at an even more acute angle relative to the basal surfaces, and the two alkyl side chains interacting directly with one basal surface each (Figure 10). This configuration in K-SWy-2 appears to be at least partially stabilized by H-bonds between amino hydrogens and the smectite basal oxygen atoms; two such bonds of 2.5 and 2.6 A formed between one amino group and one basal surface and were then stable for the remainder of the simulation. The rest of the atrazine molecule is forced to interact with interlayer water molecules. As the amount of interlayer water increases, neither K+ ions nor interlayer Ca2+ ions are able to form inner-sphere complexes any longer. Apparently, when atrazine and the cations are not forced together into a severely limited space as in the monolayer case, water outcompetes atrazine for interlayer ions. 82 A B 5"“: ----- cf; N"’a%-7)'.. a let. ‘— .‘ P.J. , r .1}: e i“ it'- 1‘1 ’13:» if! 2" 1.3, “‘ w .n r E '_ “Mow-.. -----.|.v- ”Pkg! ----- J":- IA— 3": Figure 10. Snapshots of molecular dynamic simulations of atrazine in the interlayer regions of two different homoionic smectites, (A) Ca-SAz-l, and (B) K-SWy-Z. Clay layers were parallel to the page and have been removed in order to expose the interlayer complexes. The atom coloring scheme is K = green, 0 = red, H = white, C =gray, N =b1ue, and C1 = green (associated with the atrazine molecule). In summary, these simulations reinforce and illuminate our emerging understanding of the interactions between smectite clays and relatively nonpolar organic solutes. New X-ray diffraction evidence (Chappell et al., 2005) suggests that the strongest sorption of atrazine to smectites occurs for interlayers containing a monolayer of water. The molecular simulations of atrazine in these monolayer systems indicate that interactions of atrazine with the clay mineral surface are by full molecular contact with both basal surfaces, thus removing most of the atrazine molecule from water. It is reasonable that the entropy increase on removing atrazine from a fully hydrated state in 83 the bulk solution to a lesser hydrated state in the smectite interlayer contributes a favorable free energy change to the system (Li et al., 2004). This effect is similar to capillary condensation of hydrophobic solutes (Corley et al., 1996), in which the free energy of sorption becomes more favorable as pores in the adsorbent become smaller. In this case, the smectite interlayer is a slit-pore and the optimal thickness for adsorption of hydrophobic molecules is approximately the molecular thickness, which corresponds (for aromatic molecules) to d-spacings near 12.5 A. Therefore, the difference in hydration status of atrazine in the interlayers is believed to account for the dramatic effect of crystalline swelling on the affinity of smectites for atrazine (Chappell et al., 2005). Another way of stating this is that the monolayer is an interlayer phase (Laird and Shang, 1997) that is much more conducive to hydrophobic organic solute partitioning from water than are other, more hydrated and expanded, interlayer phases. In a monolayer, water interactions with cations and two clay basal surfaces may reduce the entropy of water so significantly (Sposito and Prost, 1982) that interlayer water becomes almost a nonaqueous phase and the penalty for replacing it with a nonpolar organic solute vanishes. This may explain why our simulations showed that cation-organic complexes formed so much more readily in a monolayer of water than in the more hydrated interlayers. Such interlayer cation-organic complexes have been demonstrated spectroscopically for more polar solutes such as nitrobenzenes (Johnston et al., 2001; Johnston et al., 2002) and carbamates (manuscript in preparation), but not yet for atrazine. As hypothesized (Laird, 1996), the simulated atrazine tended to gravitate to uncharged and therefore relatively hydrophobic portions of the smectite basal surfaces. 84 This is in accord with observations that, all else being equal, lower-charged smectites will adsorb more atrazine (Laird, 1996; Laird et al., 1992; Sheng et al., 2001; Chapter 2) because they have more of the hydrophobic nanosites on their surfaces and also fewer interlayer cations competing for adsorption domains in the interlayer region (Sheng et al., 2002). Even in wetter systems swollen with two or three layers of water, our simulations showed that the alkyl tails of the atrazine molecule tend to interact with the uncharged nanosites on the smectite basal surfaces, i.e., valence-satisfied basal oxygens located between charge sites. 85 LITERATURE CITED Bailey, G.W., J .L. White, and T. Rothberg. 1968. Adsorption of organic herbicides by montmorillonite: Role of pH and chemical character of adsorbate. Proceedings of the Soil Science Society of America 32:222-234. Barriuso, E., D.A. Laird, W.C. Koskinen, and RH. Dowdy. 1994. Atrazine desorption from smectites. Soil Science Society of America Journal 58:1632-1638. Belluck, D.A., S.L. Benjamin, and T. Dawson. 1991. Groundwater contamination by atrazine and its metabolites: risk assessment, policy and legal implications, In L. Somasundaram and J. R. Coats, eds. Pesticide transformation products: fate and significance in the environment. American Chemical Society, Washington, DC. Biradar, DP, and AL. Rayburn. 1995. Chromosomal Damage-Induced By Herbicide Contamination At Concentrations Observed in Public Water-Supplies. Journal of Environmental Quality 24:1222—1225. Blanchard, PE, and W.W. Donald. 1997. Herbicide contamination of groundwater beneath claypan soils in North-Central Missouri. Journal of Environmental Quality 26:1612-1621. Boyd, S.A., G.Y. Sheng, B.J. Teppen, and C.J. Johnston. 2001. Mechanisms for the adsorption of substituted nitrobenzenes by smectite clays. Environmental Science & Technology 35:4227-4234. Brigatti, M.F., and S. Guggenheim. 2002. Mica crystal chemistry and the influence of pressure, temperature, and solid solution on atomistic models. Micas: Crystal Chemistry and Metamorphic Petrology 46: 1-97. Burkart, M.R., D.W. Kolpin, and DE. James. 1999. Assessing groundwater vulnerability to agrichemical contamination in the Midwest. Water Science and Technology 392103-112. Chappell, M.A., D.A. Laird, M.L. Thompson, H. Li, B.J. Teppen, V. Aggarwal, C.T. Johnston, and SA. Boyd. 2005. Influence of smectite hydration and swelling on atrazine sorption behavior. Environmental Science & Technology 39:3150-3156. Chien, Y.Y., and W.F. Bleam. 1997. Fluorine-19 nuclear magnetic resonance study of atrazine in humic and sodium dodecyl sulfate micelles swollen by polar and nonpolar solvents. Langmuir 13:5283-5288. Clay, S.A., D.E. Clay, W.C. Koskinen, and R.K. Berg. 1998. Application method: impacts on atrazine and alachlor movement, weed control, and corn yield in three tillage systems. Soil & Tillage Research 48:215-224. 86 Corley, T.L., J. Farrell, B. Hong, and M.H. Conklin. 1996. VOC accumulation and pore filling in unsaturated porous media. Environmental Science & Technology 30:2884-2891. Cruz, M., J .L. White, and J .D. Russell. 1968. Montmorillonite-s-triazine interactions. Israel Journal of Chemistry 62315-323. Curtiss, G.P., P.V. Roberts, and M.V. Reinhard. 1986. A natural gradient experiment on solute transport in a sand aquifer. 4. Sorption of organic solutes and its influence on mobility. Water Resources Research 2232059. Dao, TH, and TL. Lavy. 1978. Atrazine adsorption on soil as influenced by temperature, moisture content and electrolyte concentration. Weed Science 26:303-308. Evangelou, VP. 1998. Environmental soil and water chemistry : principles and applications John Wiley & Sons, New York. Frissel, M.J., and G.H. Bolt. 1962. Interaction between certain ionizable organic compounds (herbicides) and clay minerals. Soil Science 94:284-291. Hance, R.J. 1967. Relationship between partition data and the adsorption of some herbicides by soils. Nature 214:630-631. Israelachvili, J .N. 1992. Intermolecular and surface forces. 2nd ed. Academic Press. San Diego, CA. Jaynes, W.F., and SA. Boyd. 1991. Hydrophobicity of siloxane surfaces in smectites as revealed by aromatic hydrocarbon adsorption from water. Clays and Clay Minerals 39:428-436. Johnston, C.T., M.F. De Oliveira, B.J. Teppen, G.Y. Sheng, and SA. Boyd. 2001. Spectroscopic study of nitroaromatic-smectite sorption mechanisms. Environmental Science & Technology 35:4767-4772. Johnston, C.T., G. Sheng, B.J. Teppen, S.A. Boyd, and M.F. De Oliveira. 2002. Spectroscopic study of dinitrophenol herbicide sorption on smectite. Environmental Science & Technology 36:5067-5074. Karasawa, N., and WA. Goddard, ID. 1989. Acceleration of convergence for lattice sums. Journal of Physical Chemistry 93:7320-7327. Kolpin, D.W., J .E. Barbash, and R.J. Gilliom. 1998. Occurrence of pesticides in shallow groundwater of the United States: Initial results from the National Water-Quality Assessment Program. Environmental Science & Technology 32:558-566. 87 Laird, DA 1996. Interactions between atrazine and smectite surfaces, p. 86-100, In M. T. Meyer and E. M. Thurman, eds. Herbicide Metabolites in Surface Water and Groundwater, Vol. 630. American Chemical Society, Washington, DC. Laird, D.A., and C. Shang. 1997. Relationship between cation exchange selectivity and crystalline swelling in expanding 2:1 phyllosilicates. Clays and Clay Minerals 45:681-689. Laird, D.A., and PD. Fleming. 1999. Mechanisms for adsorption of organic bases on hydrated smectite surfaces. Environmental Toxicology and Chemistry 18:1668- 1672. Laird, D.A., and BL. Sawhney. 2002. Reactions of pesticides with soil minerals, p. 765— 793, In J. B. Dixon and D. Schulze, eds. Soil mineralogy with environmental applications, Vol. SSSA Book Ser. 7, 2nd ed. Soil Science Society of America, Madison, WI. Laird, D.A., E. Barriuso, R.H. Dowdy, and W.C. Koskinen. 1992. Adsorption of atrazine on smectites. Soil Science Society of America Journal 56:62-67. Li, H., B.J. Teppen, C.T. Johnston, and SA. Boyd. 2004. Thermodynamics of nitroaromatic compound adsorption from water by smectite clay. Environmental Science & Technology 38:5433-5442. Li, H., G.Y. Sheng, B.J. Teppen, C.T. Johnston, and SA. Boyd. 2003. Sorption and desorption of pesticides by clay minerals and humic acid-clay complexes. Soil Science Society of America Journal 67:122-131. Mader, B.T., K. UweGoss, and SJ. Eisenreich. 1997. Sorption of nonionic, hydrophobic organic chemicals to mineral surfaces. Environmental Science & Technology 31:1079-1086. Maple, J .R., M.-J. Hwang, T.P. Stockfish, U. Dinur, M. Waldman, C.S. Ewig, and AT Hagler. 1994. Derivation of Class H force fields. 1. Methodology and quantum force field for the alkyl functional group and alkane molecules. Journal of Computational Chemistry 15: 162-1 82. Mooney, R.W., A.G. Keenan, and LA. Wood. 1952. Adsorption of water vapor by montmorillonite. H. Effect of exchangeable ions and lattice swelling as measured by X-ray diffraction. Journal of the American Chemical Society 74: 1371-1374. Accelrys, Inc. 2000. Cerius2 Molecular Simulation Software, Release 4.2. Accelrys, Inc., San Diego, CA. Piatt, J .J ., D.A. Backhus, P.D. Cape], and SJ. Eisenreich. 1996. Temperature-dependent sorption of naphthalene, phenanthrene and pyrone to low organic carbon aquifer sediments. Environmental Science & Technology 30:751-760. 88 Pike, DR, and EL. Knake. 1996. Field corn and soybean pesticide use summaries. 1991-1994 National agricultural statistical service data on specific products. Special Report 1996-02. Department of Crop Sciences, University of Illinois, Urbana-Champaign. Post, J .L. 1984. Saponite from near Ballarat, California. Clays and Clay Minerals 32:147- 153. Rebhun, M., R. Kalabo, L. Grossman, J. Manka, and C. Ravacha. 1992. Sorption of organics on clay and synthetic humic clay complexes simulating aquifer processes. Water Research 26:79-84. Ribaudo, M.O., and A. Bouzaher. 1994. Atrazine: Environmental characteristics and economics of management. microforrn A1.107:699. US. Department of Agriculture, Economic Research Service, Hemdon, VA:ERS-NASS, Washington, DC. Russell, J .D., M. Cruz, J .L. White, G.W. Bailey, W.R. Payne, J .D. Pope, and J .I. Teasley. 1968. Mode of chemical degradation of s-triazine by montmorillonite. Science (Washington DC) 160:1340-1342. Schwarzenbach, RP, and J. Westall. 1981. Transport of Non-Polar Organic-Compounds From Surface-Water to Groundwater - Laboratory Sorption Studies. Environmental Science & Technology 15:1360-1367. Sheng, G.Y., C.T. Johnston, B.J. Teppen, and SA. Boyd. 2001. Potential contributions of smectite clays and organic matter to pesticide retention in soils. Journal of Agricultural and Food Chemistry 49:2899-2907. Sheng, G.Y., C.T. Johnston, B.J. Teppen, and SA. Boyd. 2002. Adsorption of dinitrophenol herbicides from water by montmorillonites. Clays and Clay Minerals 50:25-34. i Sposito, G., and R. Prost. 1982. Structure of Water Adsorbed On Smectites. Chemical Reviews 82:553-573. Sposito, G., K.M. Holtzclaw, C. Jouany, and L. Charlet. 1983. Cation Selectivity in Sodium - Calcium, Sodium - Magnesium, and Calcium - Magnesium Exchange On Wyoming Bentonite At 298-K. Soil Science Society of America Journal 47:917-921. Sun, H. 1998. COMPASS: An ab initio force-field optimized for condensed-phase applications - Overview with details on alkane and benzene compounds. Journal of Physical Chemistry B 102:7338-7364. Tanford, C. 1980. The hydrophobic effect: Formation of micelles and biological membranes John Wiley & Sons, NY. 89 Teppen, B.J., K. Rasmussen, P.M. Bertsch, D.M. Miller, and L. Schafer. 1997. Molecular dynamics modeling of clay minerals .1. Gibbsite, kaolinite, pyrophyllite, and beidellite. Journal of Physical Chemistry B 101:1579-1587. Van Blade], R., and A. Moreale. 1974. Adsorption of fenuron and monuron (substituted- ureas) by 2 montmorillonite clays. Soil Science Society of America Journal 38:244-249. Vaz, C.M.P., S. Crestana, S.A.S. Machado, L.H. Mazo, and LA. Avaca. 1997. Adsorption isotherms for atrazine on soils measured by differential pulse polarography. Electroanalysis 9:956-958. Voice, TC, and W.J. Weber. 1983. Sorption of hydrophobic compounds by sediments, soils and suspended-solids .1. Theory and background. Water Research 1711433- 1441. Weber, J.B. 1970. Mechanisms of adsorption of s-triazines by clay colloids and factors affecting plant availability. Residue Reviews 32:93-130. Weber, J .B., P.W. Perry, and RP Upchurch. 1965. The influence of temperature and time on the adsorption of paraquat, diquat, 2,4-D, and prometone by clays, charcoal, and an anion-exchange resin. Proceedings of the Soil Science Society of America 29:678-688. Welhouse, G.J., and W.F. Bleam. 1992. NMR spectroscopic investigation of hydrogen- bonding in atrazine. Environmental Science & Technology 26:959-964. Welhouse, G.J., and W.F. Bleam. 1993. Cooperative hydrogen-bonding of atrazine. Environmental Science & Technology 27:500-505. Xu, W.Z., C.T. Johnston, P. Parker, and SF. Agnew. 2000. Infrared study of water sorption on Na-, Li-, Ca-, and Mg- exchanged (SWy-1 and SAz-l) montmorillonite. Clays and Clay Minerals 48:120-131. 90 CHAPTER IV ENHANCED SORPT ION OF TRICHLOROETHENE BY SMECTITE CLAY EXCHANGED WITH CS“ 91 Enhanced sorption of trichloroethene by smectite clay exchanged with Cs+ Abstract Trichloroethene (TCE) is one of the most common pollutants in groundwater, interestingly, Cs+ is a co-contaminant at many DOE sites contaminated with TCE. Smectite clays have large surface areas, are common in soils, have high affinities for some organic contaminants, hence the potential to influence the transport of organic pollutants entering soils and sediments. The exchangeable cations present near smectite clay surfaces can radically influence the sorption of organic pollutants by soil clays. This research was undertaken to determine the effect of Cs+, and other common interlayer cations, such as K+ and Ca2+, on the sorption of TCE by a reference smectite clay saponite. Cs-saturated clay sorbed the most TCE, up to 3500 mg/kg, while Ca—saturated smectite sorbed the least. We hypothesize that the stronger sorption of TCE by the Cs- smectite can be attributed to the lower hydration energy and hence smaller hydrated radius of Cs+, which expands the lateral clay surface domains available for sorption. Also, Cs-smectite interlayers are only one or two water layers thick, which may drive capillary condensation of TCE. Our results implicate enhanced retention of TC E in aquifer materials containing smectites accompanied by Cs+ co-contamination. 92 Introduction Chlorinated solvents such as trichloroethene (TCE, C2HC13) are common environmental contaminants near thousands of DOE, DOD, NASA and private industry facilities. TCE has been extensively used for rocket engine flushing, metal degreasing, and organic extractions due to its effectiveness, noncorrosivity, and nonflammability. In 1994 alone, 42 million pounds of TCE were released into soil and water environments in the USA, and TCE contamination was identified in at least 852 sites among the 1416 sites proposed in the US. EPA National Priority list (Doherty, 2000). TCE is also one of the pollutants detected most frequently and at high concentrations in water of industrial regions (Patrick et al., 1987), and drinking water sources tested in the US in 1997 showed a wide range of TCE contamination (Griffin, 2003). The US EPA Drinking Water Standard for TCE is 0.005 mg L'1 (Spitz and Moreno, 1996), which is just 0.0004% of its water solubility (water solubility of TCE is 1100 mg L'1 at 25°C, (Verschueren, 2001)), therefore, even small amounts of TCE leaking into a drinking water supply can pollute a large water volume. TCE is toxic, resistant to degradation and therefore persistent in subsurface environments (Li, 2001). There is a need to accurately predict the long-terrn fate and movement of TCE in soils and subsurface environments. Sorption by soils and sediments is a major process that controls the transport, bioavailability, degradation, and ultimately the fate of organic contaminants in the environment (Koskinen and Harper, 1990). Sorption of organic chemicals diminishes their bioavailability to target organisms (Alexander, 2000), affects chemical and microbiological transformations (Gamble and Khan, 1985), and reduces risk of groundwater contamination (Sonon and Schwab, 1995). Clay minerals and soil organic 93 matter (SOM) are typically the two most important components in controlling the fate of organic chemicals that enter soils and sediments. Most research conducted in the last two decades has indicated that SOM is the primary sorptive phase for the uptake of nonpolar or slightly polar organic contaminants (Chiou et al., 1979; Miltner et al., 1989; Ong and Lion, 1991; Wilmanski and Vanbreemen, 1990; Xia and Ball, 2000; Xing et al., 1996; Zimmer et al., 1989). Transport models often use soil-organic carbon-normalized sorption coefficients (K...) to predict the distribution of organic chemicals between soil and water, but this approach ignores the potential contribution of soil minerals to the retention of organic contaminants in soils and subsoils. Ofien this approach is justified on the basis that water molecules can effectively compete with organic compounds for the sorption sites on mineral surfaces, though this has proven to be an oversimplification. Clay minerals are abundant in soils and aquifer materials, have large surface areas, so have a great potential for sorption of organic compounds. Of the various clay minerals commonly found in the soil environment, smectites are most important for interacting with organics due to their high surface areas, moderate cation exchange capacities (CECS), and surface reactivities (McBride, 1994). Several previous studies have attempted to identify the critical ratio of clay minerals/SOM at which sorption by the mineral phase plays an important role (Farrell and Reinhard, 1994; Grundl and Small, 1993; Karickhoff, 1984). Karickhoff (1984) concluded that at critical ratio of swelling clays to soil organic C >30, the mineral phase in whole soil becomes important for sorption of simazine. Grundl and Small (1993) estimated the critical ratio of 62 for atrazine and 84 for alachlor at which mineral phase sorption accounted for half of the overall sorption. Farrell and Reinhard (1994) concluded that for high clay to natural 94 organic matter ratios, mineral adsorption dominates organic matter for sorption of organics. Adsorption of TCE has been studied by a number of workers, but most such studies are limited to organo-clays (Sheng et al., 1998; Zhao and Vance, 1998), zeolites (Alvarez-Cohen et al., 1993; Farrell et al., 2003; Kleineidam et al., 2002; Li, 2001), and carbon materials (Kleineidam et al., 2002; Langer et al., 1999). Only a few studies on the interactions of clays with volatile, hydrophobic, low molecular weight halocarbons such as TCE have been conducted to date (Estes et al., 1988; Farrell and Reinhard, 1994; Rogers and McFarlane, 1981; Singhal and Singh, 1976; Werth and Reinhard, 1997). Earlier studies have shown that smectite clays can adsorb organic solvents such as TC E (1.3 mg g'1 clay, (Farrell and Reinhard, 1994)), and tetrachloroethene (PCE) (5.4 mg g" clay, (Estes et al., 1988)), to an appreciable extent. Ong and Lion (1991) observed that under oven-dried conditions the mineral fraction of whole soil plays a significant role in the sorption of TCE, but competition from water reduces TCE sorption on mineral surfaces. Based on the higher sorption by a low organic carbon containing silt fraction, Pignatello (1990) concluded that the mineral fi‘action of whole soil plays an important role in TCE sorption. He also concluded that mineral fraction of whole soil has a significant role in the slow release of residual fraction as organic matter in interstitial pores of particle aggregates are protected from equilibrium with bulk solution by such mineral fractions (Pignatello, 1990). Results from the few previous studies on smectites are inconsistent. For example, Farrell and Reinhard (1994) reported that montmorillonite sorbed 1300 mg kg'1 TCE from the gas phase at 100% relative humidity, while Estes et a1. (1988) observed little or no TCE sorption by a smectite rich soil. Rogers and 95 McFarlane (1981) reported that adsorption of TC E, carbon tetrachloride, and ethylene dibromide by montmorillonite was variable, ranging from 0 tol 7 % of the chemical added to the clay, depending on the type of the clay or interlayer cation. They reported that while A13+ saturated montmorillonite sorbed 17% of the added TCE, Ca2+ saturated montmorillonite did not sorb any TCE at all. Thus, there are some indications that the sorption of chlorinated solvents by smectites is cation-dependent. One effect of changing the interlayer cations in smectites is to change the layer-spacing of the clay. Farrell et al. (2002) indicated that TCE adsorption is energetically most favorable in pores that are just large enough to accommodate a TCE molecule. The theory of capillary phase separation says that sorption of a hydrophobic molecule into a slit-pore should become more favorable in materials with small mesopores or micropores (Corley et al., 1996). We therefore hypothesize that TCE should most effectively compete with water for sorption sites in smectite clay interlayers when the smectite layer spacing is ~12.5 A, as these clay interlayers are just large enough to accommodate TCE molecules. Cs+ saturated smectites generally equilibrate with ~ 12.5 A layer spacings at 100% humidity and even in aqueous suspensions (Sato et al., 1992), indicating that there is just one layer of water between smectite layers. Thus, a priori, Cs+- smectites appear to possess optimal properties for TCE sorption. By the same reasoning, TCE should not be able to compete as effectively for sorption sites on Ca2+ saturated smectites because the strongly hydrated Ca2+ ions (AGhyd Ca2+: -1593 kJ/mol while AGhyd Cs+: -284 kJ/mol, (Evangelou, 1998)) cause greater swelling of the clay interlayers to d-spacings 218 A. Sorption sites in these larger slit-pores should not be as energetically favorable for TCE condensation from water 96 (Corley et al., 1996), so Ca2+ saturated clays are expected to sorb less TCE than C 3+ saturated clays. This hypothesis is of environmental interest, since TC E and C 5+ co-contamination has been reported in groundwater plumes at a number of DOE sites (Brigmon et al., 1998; McKinley et al., 2001; Riley and Zachara, 1992; Zachara et al., 2002). As one example, mixed liquid waste containing both Cs+ and TCE was injected through a single well into the Snake River Plain aquifer at Idaho National Engineering and Environmental Laboratory from 1953 to 1972 (National Research Council, 2000). Based on these hypotheses and observations, this study was initiated to study the adsorption of TCE by saponite, a 2:1 trioctahedral clay, saturated with different exchangeable cations (K+, Cs+ and Ca2+). K+ and Ca2+ were selected, as they are common exchangeable cations in the soil environment. Cs+ was selected primarily because TC E and Cs+ co-contamination has been reported in groundwater plumes at a number of DOE sites (Brigmon et al., 1998; McKinley et al., 2001; Riley and Zachara, 1992; Zachara et al., 2002). Secondly, Cs+ has a smaller hydrated radius compared with other cations studied here, hence is expected to create a more favorable interlayer environment (i.e. less hydrated) for TCE sorption. The negatively charged sites on saponite clay originate largely from tetrahedral substitution. 97 Materials and Methods Clays The reference clay, saponite (SapCa-2, CEC 95 cmol kg") used in the adsorption study, was purchased from the Clay Minerals Society Source Clay Repository (Department of Agronomy, Purdue University, West Lafayette, Indiana, USA). The < 2 pm clay-sized particles were separated by wet sedimentation, and then exchanged with K+, Cs+, and Ca2+ to prepare homoionic clays. To do so, 25 g of clay were dispersed in a 0.1 M chloride solution (500 ml) of the respective cation. The clay suspensions were shaken for 24 h, and fresh chloride salt solutions were used to replace the original solutions afier centrifugation. This process was repeated four times to ensure complete cation saturation. The excess chlorides were removed by repeatedly washing with Milli-Q water until CT was determined negative by reacting with AgNO3 solution. The clay suspensions were then quick-frozen, freeze-dried and stored in a closed container for later USO. Solute Analytical grade TCE used in the adsorption experiment was purchased from Aldrich Chemical Co. (Milwaukee, Wisconsin) with a purity of >99%, and used as received. The selected physicochemical properties of TCE are as follows: molecular weight 131.4 g mol", vapor pressure at 30°C is 12.5 kPa, boiling point 87 °C, density 1.45 g cm'3, log K0w at 25°C is 2.53, dipole moment 0.9 Debye, molar volume 89.87 cm3 mol'1 at 25°C, and average molecular diameter is 6.8 A (Farrell and Reinhard, 1994). 98 Determination of Henry ’3 Law Constant The equilibrium headspace technique was used to determine Henry’s Law constant of TCE in different electrolytes, which is similar to the method reported by Garbarini and Lion (1985) and Lincoff and Gossett (1984). Determinations of Henry’s Law constant as well as sorption isotherms were conducted using 20 mL capacity glass headspace GC-vials. Bottle volume was determined by weight by filling to capacity with water at room temperature (23 i 2 °C). Three different TCE concentrations were prepared in 0.1 M KCl and CsCl, and 0.05 M CaClz. Same volume (25 pL) of stock TCE solution (prepared immediately prior to their use) (500 mg L'l) was injected into the vials containing different volumes of each electrolyte solution to create different concentrations. Vials were capped immediately with Teflon-lined silicone septa and aluminum crimp caps to minimize volatile loss of TCE. The vials were then shaken reciprocally overnight at 40 rpm at room temperature (23 i 2 °C); preliminary studies showed that equilibrium was achieved within this time-period. TCE concentration in the headspace of each vial was analyzed using a Varian 3700 gas chromatograph with flame ionization detector (GC-FID) connected with a HP 19395A Headspace Autosampler. The dimensionless Henry’s Law constant (Hg) was calculated (Lincoff and Gossett, 1984) as: CgZVIZ —CglVIl (1) ‘=C V —Cg2V gl :1 82 where C g. and C 32 are the TCE concentrations in gas phase of vials 1 and 2, Vgl and V£2 are the gas volumes in vials l and 2, and V,l and V,2 are the solution volumes in vials l 99 and 2. The experiment was performed in triplicate, and at least duplicate results are presented. We used these values of Henry’s Law constant to calculate the distribution of TCE in the sorption experiments, as Henry’s Law was dependent on the cation in solution. Sorption Isotherm Measurement TCE sorption isotherms were determined using a batch equilibrium headspace technique (Garbarini and Lion, 1985). Freeze-dried clay (0.1 to 0.6 g) was suspended in 5.0 mL of electrolyte solution (0.1 M for CsCl and KC], and 0.05 M for CaClz). A series of initial TCE concentrations (0.5, 1.0, 2.0, 5.0, 10, 20, 40, 70 and 100 mg L") were prepared by injecting different volumes of stock TCE solution (500 mg L") (prepared immediately prior to their use in respective electrolyte solution) into vials using a syringe. Vials were capped immediately with Teflon-lined silicone septa and aluminum crimp caps to minimize volatile losses of TCE. The vials were then shaken reciprocally overnight at 40 rpm at room temperature (23 i 2 °C); preliminary studies showed that equilibrium was achieved within this time. TCE concentration in headspace of the vials was analyzed using a Varian 3700 GC-FID as above. All isotherm experiments were performed in triplicate, and at least duplicate results are presented. After reaching partitioning equilibria, TCE distributions in air, solution and clay were determined by comparing two systems, viz. systems containing clay and systems without clay. The same solution volume (V, ), gas volume (Vg ), total mass of TC E ( M ,.) was applied to both systems. Under such circumstances, sorption equilibria were evaluated using the following relationships: 100 a. System without clay (at equilibrium): MT =C,,V,+Cg,Vg (2) b. System with clay (at equilibrium): MT = CHI/I +Cg2Vg + C Molar (3) day where C ,, and C ,2 are the equilibrium TCE concentrations in solution phase of systems without and with clay, respectively, Cg1 and C g 2 are equilibrium TCE concentrations in the gas phase of systems without and with clay, respectively, C is the TCE clay concentration in the clay phase, and M is the mass of clay (g) added to the system. clay Since the same amounts of TCE were added into both systems, equations (2) and (3) are equal, thus: Cll V! + Cgl Vg = C12 V! + CgZVg + CclayMclay (4) From Henry’s Law, we know that: C8 C1 = H— (5) and we determined Hc for each electrolyte using equation (1) as described above. Substituting C, from equation (5) into equation (4), we get: M clay (6) day Cg'V CV—CgZV C V C F14» glaz-H 1+ ng+ (' t'. 101 From equation (6) the amount of TC E sorbed on the clay (Cclay) can be determined as: C —C.. * V/H V C.,.,..=( g‘ ”MK’ ‘)+ g] (7) (lay C 3, and C g 2 are determined using a Varian 3700 GC-FID. The TCE concentrations in aqueous solution and clay phases were calculated using equation 5 and 7, respectively. X -ray Diffraction Clays were mixed with water to form suspensions, which were used to prepare clay films for X-ray diffraction analysis. The suspensions were dropped onto glass slides and air-dried overnight at ambient conditions to obtain oriented clay films. The clay films were then allowed to equilibrate with water vapor for 2 d in a closed container containing distilled water to expose the clay films at 100% relative humidity. X-ray diffraction patterns were recorded using Cu-Ka radiation and a Philips APD 3720 automated X-ray diffractionmeter, fitted with an APD 3521 goniometer and a 0-compensating slit, a 0.2 mm receiving slit and a diffracted beam graphite monochromator. Diffraction patterns were recorded from 5.5 to 16.0 °20 at a step of 0.005 °20 with 4 s/step. 102 Results and Discussion Henry ’5 Law Constant Henry’s Law constant of TCE was found to be different in the three different electrolyte solutions used for the sorption study. The values of Henry’s Law constant were determined to be 0.61 for 0.1 M KCl, 0.74 for 0.05 M CaClz, and 0.92 for 0.1 M CsCl, which are all different from the value (0.40) in the pure water-air system (Verschueren, 2001). The presence of salt in the water phase makes the aqueous solution more unfavorable for the dissolution of TCE; thereby enhancing the Henry’s Law constant. Since these differences are large and since the Henry’s Law constant plays a significant role in the calculation of sorption (eqs 5 and 7), we measured these values of Henry’s Law constant for each salt solution and used these values of Henry’s Law constant to calculate the TCE aqueous concentrations in the relevant electrolyte solutions in our sorption experiments. Sorption Isotherm TCE sorption isotherms are shown in Figure 1. The data points are the average values of equilibrium concentration (X-axis) and amount of TCE sorbed (Y-axis), and the error bars show the standard deviations. Ca-saturated saponite sorbed the least TC E (Figure 1a) and Cs-saturated saponite sorbed the most TCE (Figure 1b). K-saturated saponite sorbed roughly twice that amount of the Ca-saponite (Figure 1a). The sorption of 0.34% (wt/wt.) TCE by the Cs-smectite at an equilibrium concentration of 90 mg L'I is the largest TCE sorption reported for any clay mineral that we know of, though Alvarez- Cohen et a1. (1993) observed 20% TCE sorption for a low-charge zeolite, and Farrell et 103 al. (2003) observed up to 50% TCE sorption for another low-charge zeolite under both dry air and 100% relative humidity, while Farrell and Reinhard (1994) reported 0.13% (wt/wt.) TCE sorption by montmorillonite and Rogers and McFarlane (1981) did not observe any TCE sorption by Ca2+ saturated montmorillonite. While solute transport models generally assume no TC E sorption to clays in aquifer materials, Figure 1 shows that there might be significant sorption of TCE to aquifer smectites especially if Cs+ is a co-contaminant in the plume. 104 800 Amount of TCE sorbed, mg/kg 100 4000 3000 2000 1000 0 , I 1 r O 20 40 60 80 100 Equilibrium Concentration of TCE, mg/L Figure 1. Sorption isotherms representing trichloroethene uptake from water by reference saponite (SapCa-Z) clay saturated with (a) calcium, and potassium, and (b) cesium. 105 The lower effectiveness of Ca-SapCa-Z may be due in part to the larger enthalpy of hydration of Ca2+. The larger hydration sphere around Ca2+ causes Ca—saponite to swell in water, thus expanding the slit-pore and decreasing the energy gained by removing TC E from aqueous solution (Corley et al., 1996). Also, the larger hydration sphere of C a2+ obscures a larger fraction of the lateral siloxane surface between exchangeable cations compared to those associated with either Cs+ or K”, and thereby effectively shrinks the size of the adsorptive domains for Ca2+ saturated clays. Conversely, the higher effectiveness of Cs-SapCa-2 in sorption may be due in part to the lower hydration energy of Cs+ and hence smaller hydrated radius of Cs+ compared to K+ or Ca2+. The relatively weak hydration of Cs)r also causes Cs-saponite to swell less in water, which potentially provides a more favorable sorption domain (Corley et al., 1996), and allows the organic to compete better with water for interlayer sites. The fewer water molecules associated with exchangeable Cs’, for example, obscure a smaller fraction of the clay surface than those associated with Ca2+, resulting in the increase in the adsorptive domains for Cs+ saturated clays. Zhu et al. (2004) also reported high sorption of polycyclic aromatic hydrocarbons (PAHs) by Cs-saturated clays. They hypothesized that relatively weak hydration of Cs+ resulted in stronger cation-1r interaction ensuring higher sorption of PAHs, while bivalent cations like Ca2+ having high hydration energies induced weak cation-1t interactions, resulting in low sorption. While ‘hydrated radii’ are poorly defined, typical estimates of hydrated radii are 9.6 A for Ca”, 5.3 A for K+, and 3.6 A for C 5+ (Evangelou, 1998). Thus the cross-sectional area of hydrated Ca2+ is about 290 A2, more than seven times that of Cs+ (40 A2); therefore one hydrated Ca2+ takes up 3.5 times as much room in the clay interlayer as do the two Cs+ ions that Ca2+ replaces. Using these 106 hydrated radii along with an idealized smectite basal surface area of 750 m2 g"l and the 95 cmol kg'I CEC of saponite SapCa-2, we compute that the hydrated radii of Ca2+ or K+ must overlap in the interlayer regions of homoionic SapCa-2, meaning that TCE would have to compete with strongly bound water for interlayer sorption sites, which it apparently does not do very well. On the other hand, Cs" and its hydration shells are projected to occupy only about 67% of the interlayer space, implying that TCE has to compete only with less strongly bound water in the Cs-smectite interlayers. This may explain why the Cs-smectite is a much more effective sorbent for TCE. Sheng et al. (2001; 2002) also observed similar effects of cation exchange on domain size for the sorption of atrazine and DNOC by SWy-2 clay. Of the cations studied here, Cs-smectite is most likely to have a 12.5 A layer spacing, even in aqueous suspension. If so, and assuming TCE sorption is in the interlayer, then at most a monolayer of TCE can sorb and the system should exhibit Langmuir-type behavior. The best-fitting Langmuir parameters are a monolayer coverage of about 4500 mg kg'l, and Langmuir K of about 0.032. This Langmuir function does not fit the data particularly well at low concentrations because of the upward curvature of the isotherm data, but the Freundlich curve is even worse for the same reason. Similarly, a Langmuir isotherm does not fit well for K-smectite at low concentrations, but can be fit at high concentrations (Figure 1a). The Ca-system on the other hand will never have a 12.5 A layer spacing in suspension, so sorption of a simple monolayer of TCE is unlikely in the Ca-smectite interlayer. For the Ca-system there was no evidence of an adsorption plateau, so there is no way to fit the Langmuir equation. Therefore, we report F reundlich l-N LN parameters for our Ca-system: Kr (mg kg'l), the Freundlich sorption coefficient is 107 3.11, and N (unitless) is 1.16. The N value > 1.0 for Ca-smectite is consistent with earlier results (Li et al., 2003; Sheng et al., 2001) for hydrophobic compound sorption to Ca-smectites. X -ray Diffraction Differences in the swelling behavior between Cs- and Ca—SapCa-2 may also contribute to their differential adsorption of organic compounds. Many Cs-smectites equilibrate with ~12.5 A layer spacings (hydrate monolayer structures) at 100% humidity and even in aqueous suspension (Sato et al., 1992). All Ca-smectites, on the other hand, swell to more than 15 A in suspension. Our results show that Cs-saturated saponite swells to 12.1 A at 100% humidity (K-saturated saponite swells to 12.2 A), while Ca-saturated saponite swells to 14.9 A at 100% humidity (Figure 2). Of course, the d-spacings of these clay films may be smaller than those of the smectites in aqueous suspension when the adsorption experiments were carried out, although the clay films were exposed to 100% RH. Teppen et al. (1998) concluded that, for idealized siloxane surfaces of layer charge zero, the most stable mode of TCE sorption occurs by full molecular contact, coplanar with the clay surface. They further concluded that this kind of interaction is suppressed by increasing water loads. The average molecular diameter of TCE is 6.8 A (Farrell and Reinhard, 1994) and the thickness perpendicular to the molecular plane is only about half that. We propose that in a system with a monolayer of water (K+, and Cs+-saturated clays near 12-A d-spacings as in Figure 2), TCE intercalates into clay interlayers and, coplanar with the clay basal surfaces, interacts with two opposing clay layers by fiill molecular contact with both via van der Waals and electrostatic forces. That is, TCE like many 108 planar aromatic molecules is only about 3 to 3.5 A thick so fits very well between two smectite layers at d-spacings of 12-12.5 A (Sheng et al., 2002). Thus Cs-smectite, which can reasonably be expected to have a 12.5-A d-spacing even in aqueous suspension (Berend et al., 1995; Kawano and Tomita, 1991; MacEwan and Wilson, 1980; Suquet and Pezerat, 1987; Suquet et al., 1975; Suquet et al., 1977), may have the optimal slit- pore to maximize the energy of TCE condensation from solution (Corley et al., 1996), because if the interlayer region was any more narrow than 12 A, then TCE could no longer intercalate at all because it simply could not fit. Intensity .- d .1 2 Theta Figure 2. X-ray diffraction pattern of reference saponite (SapCa-Z) clay in oriented clay films as a function of different cation saturation at 100% humidity. 109 Given a 3-A thick interlayer region, 33% of the basal surface not covered by Cs+ hydration shells as described above, and the TCE molar volume cited above, we predict that the Cs-saponite might be able to sorb up to 10% TCE by weight if all interlayer volume outside primary Cs+ hydration shells were filled with TCE. Since we observed only 0.34% sorption (Figure 1), our assumptions may be simplistic in that a) The kinetics of tortuous TCE diffusion to many interlayer sites may be slow, or b) TCE is not able to displace other interlayer water outside primary Cs+ hydration shells. In this regard, a tetrahedrally substituted clay may have been an inopportune choice in that tetrahedral charge is much more localized on fewer surface oxygens (McBride, 1989) and exchangeable cation ordering is only found in tetrahedrally substituted smectites (Odom, 1984). The exchangeable cations (at a given hydration state) can always approach a tetrahedral site more closely than an octahedral site (Bleam, 1990), and thereby strengthen the hydrogen bonding of interlayer water to the siloxane surface itself (Yariv, 1992) Sorption isotherms for Cs-, K-, and Ca-saturated saponite are nonlinear (Figure l a and b). This nonlinearity could be due to several possible causes: 1) The heterogeneous distribution of sorption sites on the clays, 2) Equilibrium was not achieved during the course of the study, or 3) At least some of the postulates of a Langmuir-type sorption are satisfied, particularly the presence of a finite number of sorptive sites that, when fully occupied, constitute a ‘monolayer’ of adsorbate coverage. Li and Werth (2001) concluded that since sorbate molecules first occupy those sorption sites that have the highest adsorption potential, therefore sorption isotherms of TCE on two different natural solids were nonlinear. Similarly, Estes et al. (1988) and Hamaker and Thompson (1972) 110 did not observe equilibrium for sorption of organic on montmorillonite clay over a period of 30 days and concluded that slow changes in clay surface that resulted in the continuous exposure of new adsorption sites were the primary reason for the failure to achieve equilibrium. If our hypotheses about TCE sorption to the Cs-saponite are correct, then the smectite interlayer should behave very nearly like an ideal Langmuir system because the 12-A d-spacing allows only monolayer coverage in the interlayer and the lateral adsorption domains between hydrated cations are only a fraction of a finite smectite basal surface area. The type of smectite clay also affects TCE sorption. For example, our results show that Ca-saturated saponite sorbed ~ 460 mg TCE/kg at the highest equilibrium concentration, but Rogers and McFarlane (1981) in a study with Ca-saturated montmorillonite (dioctahedral, CEC 80 cmol/kg) did not observe any apparent sorption of TCE. Two possible reasons for these differences are: (i) differences in the CEC of the clays, and (ii) type of the clay. However, we can ignore the differences in CEC to be responsible for the differences in sorption of TCE by these two clays. Generally, the lower the CEC of clay, the smaller is the clay surface occupied by the exchangeable cations, which makes more clay siloxane surface area available for sorption. Several workers have reported an inverse relationship between the CEC of the clay and the amount of the organic adsorbed (Laird et al., 1992; Lee et al., 1990; Sheng et al., 2001; Sheng et al., 2002). Based on these results, Ca-saturated montmorillonite of Rogers and McFarlane (1981) should have sorbed more TCE than that observed in the present investigation. Therefore, it seems that differences in TCE sorption may be related to the source of negative charge on the clays and not to differences in the CEC of the clays. The 111 negative charge on the saponite clay arises from isomorphous substitution in tetrahedral structures. Generally, for a given layer charge, clays with negative charge from tetrahedral substitution sorb more organics than do clays where negative charge develops from octahedral substitution. Compared to smectites with negative charges from octahedral isomorphous substitution, the 2:1 smectites with negative charges from tetrahedral isomorphous substitution exhibit less swelling of the interlayer space in water. The theory of capillary phase separation (Corley et al., 1996) thus implies that sorption of hydrophobic molecules from water should be more favorable in tetrahedrally versus octahedrally substituted smectites (for a given layer charge and interlayer cation), and such behavior has been observed (Laird et al., 1992; Sheng et al., 2001) though not explained. 112 LITERATURE CITED Alexander, M. 2000. Aging, bioavailability, and overestimation of risk from environmental pollutants. Environmental Science & Technology 34:4259-4265. Alvarez-Cohen, L., P.L. McCarty, and RV. Roberts. 1993. Sorption of trichloroethylene onto a zeolite accompanied by methanotrophic biotransforrnation. Environmental Science and Technology 27:2141-2148. Berend, I., J .M. Cases, M. Francois, J .P. Uriot, L. Michot, A. Masion, and F. Thomas. 1995. Mechanism of adsorption and desorption of water-vapor by homoionic montmorillonites .2. The Li+, Na+, K+, Rb+ and Cs+~ exchanged forms. Clays and Clay Minerals 43:324-336. Bleam, W.F. 1990: The nature of cation-substitution sites in phyllosilicates. Clays and Clay Minerals 38:527-536. Brigmon, R.L., N.C. Bell, D.L. Freedman, and OJ. Berry. 1998. Natural attenuation of trichloroethylene in rhizosphere soils at the Savannah River Site. Journal of Soil Contamination 7:433-453. Chiou, C.T., L.J. Peters, and V.H. Freed. 1979. Physical Concept of Soil-Water Equilibria For Non-Ionic Organic-Compounds. Science 206:831-832. Corley, T.L., J. Farrell, B. Hong, and M.H. Conklin. 1996. VOC accumulation and pore filling in unsaturated porous media. Environmental Science & Technology 30:2884-2891. Doherty, R.E. 2000. A history of the production and use of carbon tetrachloride, tetrachloroethylene, trichloroethylene and 1,1,1- trichloroethane in the United States: Part 2 - Trichloroethylene and 1,1,1-trichloroethane. Environmental Forensics 1:83-93. Estes, T.J., R.V. Shah, and V.L. Vilker. 1988. Adsorption of low molecular weight halocarbons by montmorillonite. Environmental Science and Technology 22:377- 381. Evangelou, VP 1998. Environmental soil and water chemistry : principles and applications John Wiley & Sons, New York. Farrell, J ., and M. Reinhard. 1994. Desorption of halogenated organics from model solids, sedimants, and soil under unsaturated conditions. 1. Isotherrns. Environmental Science and Technology 28:53-62. 113 Farrell, J ., C. Manspeaker, and J. Luo. 2003. Understanding competitive adsorption of water and trichloroethylene in a high-silica Y zeolite. Microporous and Mesoporous Materials 59:205-214. Farrell, J ., J. Luo, P. Blowers, and J. Curry. 2002. Experimental and molecular mechanics and ab initio investigation of activated adsorption and desorption of trichloroethylene in mineral micropores. Environmental Science & Technology 36:1524-1531. Gamble, D.S., and S.U. Khan. 1985. Atrazine hydrolysis in soils: Catalysis by the acidic functional groups of fulvic acid. Canadian Journal of Soil Science 65:435-443. Garbarini, DR, and L.W. Lion. 1985. Evaluation of sorptive partitioning of nonionic pollutants in closed systems by headspace analysis. Environmental Science & Technology 19:1122-1128. Griffin, BM. 2003. Microbial reductive dechlorination of chlorinated ethenes: Ecology and Application. Doctorate, Michigan State University, East Lansing. Grundl, T., and G. Small. 1993. Mineral contributions to atrazine and alachlor sorption in soil mixtures of variable organic carbon and clay content. Journal of Contaminant Hydrology 14:117-128. Hamaker, J .W., and J .M. Thompson. 1972. Adsorption, p. 49-143, In C. A. I. Goring and J. W. Hamaker, eds. Organic chemicals in the soil environment, Vol. 1. Marcel Dekker, New York. Karickhoff, SW. 1984. Organic pollutant sorption in aquatic systems. Journal of Hydraulic Engineering 110:707-735. Kawano, M., and K. Tomita. 1991. X-ray powder diffraction studies on the rehydration properties of beidellite. Clays and Clay Minerals 39:77-83. Kleineidam, S., C. Schuth, and P. Grathwohl. 2002. Solubility-normalized combined adsorption-partitioning sorption isotherms for organic pollutants. Environmental Science & Technology 36:4689-4697. Koskinen, W.C., and SS. Harper. 1990. The retention process: Mechanisms, p. 51-78, In H. H. Cheng, ed. Pesticides in the soil environment: Processes, impacts, and modeling. Soil Science Society of America, Madison, WI. Laird, D.A., E. Barriuso, R.H. Dowdy, and W.C. Koskinen. 1992. Adsorption of atrazine on smectites. Soil Science Society of America Journal 56:62-67. Langer, V.W., K.S. Novakowski, and AD. Woodbury. 1999. Sorption of trichloroethene onto stylolites. Journal of Contaminant Hydrology 40: 1-23. 114 Lee, J .-F., M.M. Mortland, C.T. Chiou, D.E. Kile, and SA. Boyd. 1990. Adsorption of benzene, toluene, and xylene by two tetrarnethylammonium-smectites having different charge densities. Clays and Clay Minerals 38:113-120. Li, C. 2001. Molecular simulation of competition between water and trichloroethene for adsorption to mineral surfaces. Masters, Michigan State University, East Lansing. Li, H., G.Y. Sheng, B.J. Teppen, C.T. Johnston, and SA. Boyd. 2003. Sorption and desorption of pesticides by clay minerals and humic acid-clay complexes. Soil Science Society of America Journal 67:122-131. Li, J ., and OJ. Werth. 2001. Evaluating competitive sorption mechanisms of volatile organic compounds in soils and sediments using polymers and zeolites. Environmental Science & Technology 35:568-574. Lincoff, A.H., and J .M. Gossett. 1984. The determination of Henry's constant for volatile organics by equilibrium partitioning in closed systems, p. 17-25, In W. Brutsaert and G. H. J irka, eds. Gas transfer at water surfaces. D. Reidel Publishing Co., Dordrecht, Holland. MacEwan, D.M.C., and M.J. Wilson. 1980. Interlayer and intercalation complexes of clay minerals, p. 197-248, In G. W. Brindley and G. Brown, eds. Crystal structures of clay minerals and their X-ray identification. Mineralogical Society, London, UK. McBride, MB. 1989. Surface chemistry of soil minerals, p. 35-88, In J. B. Dixon and S. B. Weed, eds. Minerals in soil environments, 2nd ed. Soil Science Society of America, Madison, WI. McBride, MB. 1994. Environmental chemistry of soils Oxford University Press, New York. McKinley, J .P., C.J. Zeissler, J .M. Zachara, R.J. Seme, R.M. Lindstrom, H.T. Schaef, and RD. Orr. 2001. Distribution and retention of Cs-137 in sediments at the Hanford Site, Washington. Environmental Science & Technology 35:3433-3441. Miltner, R.J., D.B. Baker, T.F. Speth, and CA. Fronk. 1989. Treatment of seasonal pesticides in surface waters. Journal American Water Works Association 81 :43- 52. National Research Council. Research needs in subsurface science, National Academy Press: Washington, DC, 2000. Odom, LE. 1984. Smectite clay minerals: properties and uses. Philos Trans Royal Soc London Ser A31 1:391-409. Ong, SK, and L.W. Lion. 1991. Effects of soil properties and moisture on the sorption of trichloroethylene vapor. Water Research 25:29-36. 115 Patrick, R., E. Ford, and J. Quarles. 1987. Groundwater contamination in the United States. 2nd ed. University of Pennsylvania Press, Philadelphia, PA. Pignatello, J .J . 1990. Slowly reversible sorption of aliphatic halocarbons in soils .2. Mechanistic aspects. Environmental Toxicology and Chemistry 9:1117-1 126. Riley, R.G., and J .M. Zachara. 1992. Chemical contaminants on DOE lands and selection of contaminant mixtures for subsurface science research. DOE/ER-0547T. Department of Energy, Office of Energy Research, Washington, DC. Rogers, RD, and J .C. McFarlane. 1981. Sorption of carbon tetrachloride, ethylene dibromide, and trichloroethylene on soil and clay. Environmental Monitoring and Assessment 1:155-162. Sato, T., T. Watanabe, and R. Otsuka. 1992. Effects of layer charge, charge location, and energy change on expansion properties of dioctahedral smectites. Clays and Clay Minerals 40: 103-1 13. Sheng, G.Y., X.R. Wang, S.N. Wu, and SA. Boyd. 1998. Enhanced sorption of organic contaminants by smectitic soils modified with a cationic surfactant. Journal of Environmental Quality 271806-814. Sheng, G.Y., C.T. Johnston, B.J. Teppen, and SA. Boyd. 2001. Potential contributions of smectite clays and organic matter to pesticide retention in soils. Journal of Agricultural and Food Chemistry 49:2899-2907. Sheng, G.Y., C.T. Johnston, B.J. Teppen, and SA. Boyd. 2002. Adsorption of dinitrophenol herbicides from water by montmorillonites. Clays and Clay Minerals 50:25-34. Singhal, J .P., and GP. Singh. 1976. Studies on adsorption and interaction of 1,2- dibromo-3-chloropr0pane .1. With montmorillonites. Journal of Agricultural and Food Chemistry 24:307-310. Sonon, LS, and AP. Schwab. 1995. Adsorption characteristics of atrazine and alachlor in Kansas soils. Weed Science 43:461-466. Spitz, K.-H., and J. Moreno. 1996. A practical guide to groundwater and solute transport modeling Wiley-Interscience, New York. Suquet, H., and H. Pezerat. 1987. Parameters influencing layer stacking in saponite and vermiculite: A review. Clays and Clay Minerals 35:353-362. Suquet, H., C. de la Calle, and H. Pezerat. 1975. Swelling and structural organization of saponite. Clays and Clay Minerals 23:1-9. 116 Suquet, H., J .T. Iiyama, H. Kodama, and H. Pezerat. 1977. Synthesis and swelling properties of saponites with increasing layer charge. Clays and Clay Minerals 25:231-242. Teppen, B.J., C.H. Yu, D.M. Miller, and L. Schafer. 1998. Molecular dynamics simulations of sorption of organic compounds at the clay mineral/aqueous solution interface. Journal of Computational Chemistry 19:144-153. Verschueren, K. 2001. Handbook of environmental data on organic chemicals. 4th ed. John Wiley & Sons, New York, NY. Werth, C.J., and M. Reinhard. 1997. Effects of temperature on trichloroethylene desorption from silica gel and natural sediments .1. Isotherms. Environmental Science & Technology 31:689-696. Wilmanski, K., and AN. Vanbreemen. 1990. Competitive adsorption of trichloroethylene and humic substances from groundwater on activated carbon. Water Research 24:773-779. Xia, GS, and WP. Ball. 2000. Polanyi-based models for the competitive sorption of low- polarity organic contaminants on a natural sorbent. Environmental Science & Technology 34:1246-1253. Xing, B.S., J .J . Pignatello, and B. Gigliotti. 1996. Competitive sorption between atrazine and other organic compounds in soils and model sorbents. Environmental Science & Technology 30:2432-2440. Yariv, S. 1992. The effect of tetrahedral substitution of Si by Al on the surface-acidity of the oxygen plane of clay-minerals. International Reviews in Physical Chemistry 11:345-375. Zachara, J .M., S.C. Smith, C.X. Liu, J .P. McKinley, R.J. Seme, and PL. Gassman. 2002. Sorption of Cs+ to micaceous subsurface sediments from the Hanford site, USA. Geochimica Et Cosmochimica Acta 66:193-211. Zhao, HT, and GE Vance. 1998. Sorption of trichloroethylene by organo-clays in the presence of humic substances. Water Research 32:3710-3716. Zhu, D., B.E. Herbert, M.A. Schlautman, E.R. Carraway, and J. Hur. 2004. Cation-pi bonding: A new perspective on the sorption of polycyclic aromatic hydrocarbons to mineral surfaces. Journal of Environmental Quality 33:1322-1330. Zimmer, G., H]. Brauch, and H. Sontheimer. 1989. Activated-Carbon Adsorption of Organic Pollutants, p. 579-596, In I. H. Suffet and P. MacCathy, eds. AC S Symposium Series, Vol. 219. American Chemical Society, New York. 117 CHAPTER V MOLECULAR SIMULATIONS TO ESTIMATE THERMODYNAMICS OF POLAR ORGANIC SORPTION TO MONTMORILLONITE 118 Molecular simulations to estimate thermodynamics of polar organic sorption to montmorillonite Introduction The list of anthropogenic organic chemicals that enter the environment expands continually. Soils and sediments are important environmental sinks for these chemicals and often contain the bulk of such chemicals after release. Hence, the potential for sorption to solid phases is a key parameter in estimating the environmental risk and fate of these chemicals, and has therefore been the subject of intensive investigation (Bailey and White, 1970; Boyd et al., 2001; Chiou, 1998; Chiou et al., 1979; Haderlein and Schwarzenbach, 1993; Karickhoff and Morris, 1985; Karickhoff et al., 1979; Sheng et al., 2002; Weissmahr et al., 1999). Soil adsorption coefficients can be determined by batch equilibrium methods, but these can be difficult and expensive procedures. Despite 40 years of experimental work, soil sorption coefficients are in many cases unavailable even for high-production-volume chemicals (Sabljic et al., 1995). Even when sorption data are available for a given compound, they pertain to only a very limited number of sorbents. It would require unfeasible financial and scientific resources to measure adsorption coefficients for all pertinent chemicals interacting with all relevant natural phases. Thus there is interest in predictive models that can estimate the potential for organic compound sorption to a given solid phase (Andersson et al., 2002; Goss and Schwarzenbach, 2001). Such models could aid our prediction of contaminant transport and fate, while potentially even pre- screening the probable fates of new compounds in order to minimize environmental risk. 119 For neutral organic compounds (NOCs), a breakthrough in predictive sorption modeling came with the realization that their sorption from water to soil organic matter (SOM) is analogous to partitioning into a bulk organic phase (Chiou et al., 1979; Karickhoff et al., 1979). Numerous approaches to estimate K0c (the soil organic C- normalized linear sorption coefficient) and therefore sorption to SOM have been published during the last 25 years, such as estimation of K0C based on water solubility (Briggs, 1981; Chiou et al., 1983; Gerstl, 1990; Karickhoff, 1981), n-octanol/water partition coefficient (Briggs, 1981; Chiou et al., 1983; Gerstl, 1990; Gerstl and Mingelgrin, 1984; Karickhoff, 1981), reversed-phase high-performance liquid chromatographic (RP-HPLC) capacity factor (Guo et al., 2002; Kanazawa, 1989; Muller and Kordel, 1996), topological indices (Gerstl, 1990; Sabljic et al., 1995), molecular parameters and linear solvation energy relationships (Gramatica et al., 2000; Huuskonen, 2003; Kanazawa, 1989; Lohninger, 1994), fragment constants and structural correction factors (Tao et al., 1999), and one-parameter linear free energy relationships (LFER) (Goss and Schwarzenbach, 2001). However, there are drawbacks with most of these approaches. For example, estimation of K0c based on water solubility and octanol/water partition coefficients can be very difficult for very hydrophobic substances (Gawlik et al., 1997). Furthermore, Goss and Schwarzenbach (2001) argued that the predictive power of one-parameter LFERs (i.e., Kow) is limited, because no single parameter can completely describe all the molecular interactions that determine environmental phase partitioning processes. For NOCs with polar functional groups, specific interactions between adsorbate and adsorbent seem most likely to impact sorption. For example, the strong sorption of 120 nitroaromatic compounds to certain smectites (swelling clay minerals) seems to be driven by complexation of interlayer cations by solute -NOz groups (Boyd et al., 2001; Johnston et al., 2001; Johnston et al., 2002). In addition to such complexation, hydrogen bonding (Laird, 1997) and hydrophobic interactions (Boyd et al., 2001; J aynes and Boyd, 1991) appear to play a role in NOC sorption to clay minerals, and evidence has been marshaled for other forces such as cation-pi bonding (Zhu et al., 2004) and electron donor-acceptor complexes (Haderlein and Schwarzenbach, 1993; Weissmahr etal., 1997). Since smectites have large surface areas and can contribute to and even dominate (Sheng et al., 2001) the sorption of many NOCs, it would be useful to be able to predict the strength of NOC sorption to smectites. To do so, however, it is apparent that many different types of molecular interactions need to be considered. Molecular modeling techniques, therefore, may be appropriate tools to estimate smectite-NOC sorption affinities. Molecular simulations can help interpret experimental observation, discriminate among competing models to explain macroscopic or spectroscopic data, integrate bulk (P,V,T) and molecular properties, and test or generate scientific hypotheses (Bleam, 1993; Cygan, 2001; Lasaga and Gibbs, 1990; Sauer and Ahlrichs, 1990; Teppen et al., 1998). In particular, classical molecular simulations (often termed molecular mechanics) that parameterize all electronic effects have been used successfully to study heterogeneous clay mineral systems (Chang et al., 1997; Cygan, 1998; Delville, 1991; Greathouse et al., 2000; Kalinichev et al., 2000; Skipper, 1998; Skipper et al., 1991; Smith, 1998; Sutton and Sposito, 2001; Teppen etal., 1997). Such simulation methods are most commonly used to probe structural properties, because it is much more difficult to accurately compute energy relations (Schafer et al., 1990). 121 Nonetheless, the methods are reaching a reasonable level of maturity where it may be fruitful to test the ability of molecular mechanics methods to compute adsorption energies. Choice of compounds Several experimental studies (Haderlein and Schwarzenbach, 1993; Haderlein et al., 1996; Weissmahr et al., 1999; Weissmahr et al., 1997) indicated that nitroaromatic compounds adsorb strongly to smectite clay minerals, especially when the smectite is saturated with K+ or Cs+ counterions. In comparing molecules of very different organic functionality, there were indications that planar compounds with polar functional groups that might complex hydrated interlayer cations were most favorably adsorbed (Sheng et al., 2001). To test and refine this concept, Boyd et a1. (2001) measured the adsorption from water of a planar series of para- and meta-substituted nitrobenzenes (SNBs) by the K-saturated Arizona smectite clay mineral SAz-l, a high-charge montmorillonite. Fourier transform infrared (FTIR) spectroscopy revealed shifts in the —NOz vibrational modes of 1,3,5-trinitrobenzene (TNB) (Boyd et al., 2001) and other SNBs (Johnston et al., 2001; Sheng et al., 2002) upon adsorption to K+-smectite, results that are consistent with the complexation of K+ by —NOz groups. Such vibrational shifts were absent or less pronounced for smectites saturated with more strongly hydrated cations (i.e., Na+, M g”, Ca2+, Bar"). Molecular dynamic simulations of dinitrobenzene systems (Boyd et al., 2001; Sheng et al., 2002) predicted that each -NOz group interacted with one or more different interlayer K+ counterions. The experimental adsorption data (Boyd et al., 2001) 122 indicated that, among the nitrobenzenes tested, F reundlich complexation constants were strongest for the carboxylic methyl ester functional group, and decreased in the order: R-COOCH3 > R-NOz > R-COCH; > R-CN > R-CHO > R-OCOCH3 > R-OCH3 > R-NHg, where R = nitrobenzene. Such an ordering is plausible from a complexation standpoint, considering both the presence and steric availability of atoms that can interact favorably with KL Two hypotheses that have been presented (Boyd et al., 2001) to explain this selectivity sequence for nitroaromatic molecules are 1) Additive complexation interactions between organic molecule functional groups and multiple interlayer cations, and 2) Hydration energy differences between the various organic molecules. Recently, enthalpies were measured (Li et al., 2004) for the adsorption of three nitroaromatic solutes to smectites. Thus, the nitroaromatic compounds are a promising system for developing molecular simulation methods to study adsorption energies, because systematic bulk adsorption data (Boyd et al., 2001), spectroscopic (Johnston et al., 2001) and X-ray diffraction (Li et al., 2004; Sheng et al., 2002) investigations of interlayer environments, and adsorption enthalpy data (Li et al., 2004) are all available. A simulation model is of potential importance since it might enable us to estimate the adsorption energies of new chemicals even before they are synthesized. If we are successful, it might be possible in the future to screen interesting chemicals (i.e., pesticides) for their soil sorption and leaching potentials even before chemical synthesis and biological efficacy testing are done. Our goal in the present study is to develop and test two methods that use molecular dynamic simulations to estimate the enthalpies of adsorption for several NOCs interacting with smectites, and then correlate these enthalpy predictions with 123 experimental enthalpies and adsorption coefficients. The objectives of this study are to simulate a series of these nitroaromatic molecules in the interlayer region of montmorillonite clay, to use the molecular dynamics results to compute the relative adsorption energies for these nitroaromatics, and to determine the dominant components of the adsorption energies. Materials and Methods A model for K-SAz-l was created and molecular dynamics simulations of para- substituted nitrobenzene-clay complexes were performed in order to estimate the adsorption energies for a series of nitroaromatic compounds. In this approach, electrons are not treated explicitly, but their effects are parameterized into classical (that is, non- quantum) mechanical energy functions comprising a force field. The clay model had the composition Kl2(Al36Mg12)Si96024o(OH)4g with a CEC of 132 cmol/kg. The smectite surfaces were constructed using methods previously detailed (Teppen et al., 1997), in which a muscovite mica layer was the starting point and octahedral Mg-for-Al substitutions were random except that no two Mg octahedra were allowed to be immediate neighbors. In order to prepare for molecular dynamics simulations of nitrobenzene-clay complexes, the interlayer region of the clay was expanded to 20 A and one molecule from the series of substituted nitrobenzene was added to the unit cell, with its molecular plane perpendicular to the basal planes of the clay in an attempt to avoid biasing the final structure. The simulated loading rate was equivalent to 18 mg m'troaromatic per g clay. Water molecules were also added (in random interlayer positions) so that constant-pressure simulations resulted in 12.3 A clay layer spacings. If 124 the simulation did not result in a clay layer spacing of 12.3 A, then water molecules were either removed or added (after expanding the clay) so that the equilibrium clay layer spacing was 12.3 :1: 0.1 A, in accord with the experimental data for substituted nitrobenzenes in air-dried films of K-SAz-l. The energies were computed using a force field developed for clays (Teppen et al., 1997) combined with the polymer consistent force field (PCFF) (Maple et al., 1994; MSI, 2000) for the organics and water. As a further modification, charges on the organic solute were assigned using the COMPASS force field (Sun, 1998), which optimized organic atomic charges for condensed phase systems. In previous work (Boyd et al., 2001), we computed quantum mechanical charges for a series of nitroaromatic solutes and found COMPASS charges to more faithfully reproduce the quantum charge estimates than did PCF F charges. The charges on organic solute atoms are the most important force field parameters in determining interactions with the clay surface, interlayer water, and aqueous cations. In a typical force field, the nuclear charge and the local electron density around a given atom are subsumed into one atom-centered partial atomic charge, and the COMPASS charges for nitroaromatic solutes are illustrated in Figure 1. Commercial software (M81, 2000) was used to compute the classical molecular dynamics. No constraints were imposed on our systems during their molecular dynamics except for the contents of the simulation cell: All atoms were completely free to move, and unit cell shape and volume were unconstrained. 125 01\ /.-01 Charge N] I C1 +0.2390 H,\ /C1\ H] C2 -0.1268 C2 \ Cz/ H Nl +0.6170 /C2 / C2 H1 +0.1268 H, \C/ \H1 1 | 0. -0.4280 /N1‘ 0] '0l Figure 1. Charges on para-dinitrobenzene molecule used in simulation study. The molecular dynamics simulations were run in the NPT ensemble for 50 ps, much longer than the time required for the system volume and energy to equilibrate. Trajectories were saved from each molecular dynamics simulation, and snapshots from the post-equilibrium portion of each dynamic trajectory were used to estimate clay- organic interaction energies by computing the difference between the total energy of the system and the energies of the hydrated clay and the organic separately. A scripting program was written to automate the energy calculation for many snapshots of a given system. We used two different algorithms for converting simulation energies to predictions for experimental adsorption enthalpies. 126 Computational algorithm 1 The typical experimental scheme for measuring adsorption of aqueous organics to clays is Clay(aq) + 0rg(aq) 1::> clay-org(aq) + H20(aq) [1] Achpt where chp. is the experimentally determined adsorption energy (for example, the overall free energy, enthalpy, or entropy of the transfer process). The experimental reaction is too difficult to simulate directly, since diffusion kinetics are slower than the ns timescales that are accessible to molecular simulations. Instead, we propose an alternative path that is more computationally feasible and follows the spirit of many schemes that have been used to simulate the binding of drugs to polypeptides (Gohlke and Klebe, 2002; Halperin et al., 2002). We employ the following thermodynamic cycle for computing the energy of adsorption of an aqueous organic molecule to a hydrated clay mineral: AHexpt AH 3 AH 5 [2] Clay-cavity..." + orgmq) + H20 E:> clay-org(aq) + cavity(aq) + H20 AH4 Since the free energy, enthalpy, and entropy are state functions (Atkins, 1998), their magnitudes should be independent of path, so that AHCXPI = AH3 + AH4 + AH5 [3] 127 This thermodynamic cycle is very similar in spirit to a recent study by Goss and Schwarzenbach (2001) where the free energy change AGi for partitioning of a compound between two phases is decomposed to enable a more mechanistic understanding of the partition process (Figure 2). Partitimingofsnnllorganicsbetween aclayplnseandanaqueomphase AQ=AH3+AILclay+AHS'AHi-aqmis Figure 2. Schematic for deconposing the energy changes for tramfer ofacorrpourxifiomaqueous solutiontotheclaylayer (after Goss and Schwarzenbach, 2001). 128 Similarly, Scmidt (Schmidt et al., 2002) proposed that the free energy of transfer of a molecule between two bulk phases 1 and 2, ALZG, depends on the cohesive energy of both phases and the interactions of the molecule with both surrounding phases (van der Waals and H-bond interactions). All these interactions need to be taken into consideration when comparing the partitioning of a set of molecules between various organic phases and water. The transfer of a molecule into a bulk phase requires a cavity to be formed to accommodate the molecule (and therefore depends on the size of the molecule). For the formation of the cavity, the intermolecular interactions of the bulk phase molecules are interrupted. Thus, the stronger the cohesive interactions in the bulk phase, the more energy is required to form a cavity in the phase. To relate AH 4 of our thermodynamic cycle to Figure 2, we recognize that AH 4 = AH i-clay _ AH i-aqueous = AH (clay-cavity —> clay-org) + AH (org(aq) -—) water cavitytaql) [4] That is, AH4 is the sum of the energy changes for transferring the organic compound out of a cavity in the aqueous solution and into a cavity in the clay interlayer region. We estimated each energy term in equation 4 by equilibrating the entire organic-water (clay) system and then calculating interaction energies between subsets of the system. For example, we used the following algorithm to calculate AHimy (the interaction energy between organic compound 1' and its environment in the hydrated smectite interlayer) of Figure 2: AH...21y = AH — AH 1 — AH 2; where [5] to! 129 0 AH to, = Total clay—organic system energy 0 AHl Energy of system after organic has been deleted 0 AH2 = Energy of the system after clay layers, interlayer cations, and water have been deleted 0 Compute AHMW for many (2 150) snapshots to compile an estimate for the ensemble average value Using this same general algorithm, we also computed AHi.aqueous (the interaction energy between the organic compound and its environment in aqueous solution), AH 3 (the interaction energy between the clay phase and the water molecules taken from the clay interlayer to create a cavity the size of the organic compound: AH3 was calculated by removing water molecules from the smectite with 12.3 A layer spacing but no interlayer organic compound — see Figure 2), and AH; (the energy required to fill the cavity in water that previously held the organic compound — see Figure 2). Once the four energy terms AH..clay, AH.-aqueous, AH3, and AH5 were computed, we were able to estimate AH...pl as illustrated in cycle [2]. Note that we have neglected two necessary terms in the total energy. Both AH 3 and AH5 can be though of as containing two separate components (Gallicchio et al., 2000), one for creating a cavity in water and a second component for the interaction of the cavity contents with its environment. Calculation of the first component, cavitation, for AH3 presents some special problems: Most studies (Gallicchio et al., 2000; Hdfinger and Zerbetto, 2003; Sandberg et al., 2002) of cavitation in water work with bulk aqueous phases that can expand to accommodate the cavity without any fundamental change to 130 the phase at distance from the cavity. In contrast, smectite clays might not be able to add a cavity to their interlayer volume without swelling and thus becoming a new phase. In our molecular dynamics methods, we use periodic boundary conditions to allow a relatively small unit cell (here, just 700 atoms) represent a condensed phase by replicating the unit cell to infinity. In such a case, the composition of the unit cell is typically fixed and water cannot leave the interlayer region (for example, in response to a growing cavity). In principle, we could do a realistic simulation by constructing a reasonable clay fragment consisting of multiple smectite layers embedded in a much larger unit cell containing an aqueous solution. Then, as a cavity grew, water could actually leave the interlayer region and enter the bulk aqueous phase. We leave such a simulation for the future, as we anticipate very slow simulations due to a huge number of atoms and also slow kinetics of water diffusion out of the clay. Indeed, with such a system one might have better luck actually trying to get an organic solute to adsorb into or desorb from the interlayer region and track the enthalpy changes directly through time. For the present study, we will use literature estimates for the cavitation components of both AH3, and AH5, and simulate just the water addition or removal from the cavity. The molar volume of solid 1,4-dinitrobenzene is 103.5 cm3/mol (Lide, 1995), implying a molecular volume of 172 A3/molecule, which should be a lower limit for the molecular volume in any other phase since crystalline 1,4-dinitrobenzene packing should be maximally efficient. This is about six times the molecular volume of liquid water (30 A3/molecule), so any cavity for insertion of 1,4-dinitrobenzene into another phase should require removal of more than six water molecules. Our clay-water system (containing no 1,4-dinitrobenzene) that gave clay layer spacing of 12.3 A contained 55 water molecules 131 per unit cell while our clay-1,4-dinitrobenzene-water system of the same dam-spacing required 45. This difference of 10 water molecules per 1,4-dinitrobenzene is reasonable given packing inefficiency, and was adopted as an empirical estimate for the number of water molecules that must be removed in order to make a cavity for the nitroaromatic solutes here. Values of H3 (equation 2) were computed by removing 10 continuous water molecules out of the 55 originally present in the clay-water system (containing no organic) with deal-spacing of 12.3 A (same clay layer spacing as our clay-organic-water system), while the values of H5 (equation 2) were computed by removing 10 waters out of the 252 originally present in a periodic box of water (containing no clay or organic) (see Figure 2). The same estimates for H3 and H5 were used for each compound. Computational algorithm 2 The overall enthalpy change for adsorption of NACs to mineral surfaces from an aqueous phase can be expressed by combining equation 3 and 4 to give: AI‘iexpt : A1"I3cav + AH3int + AI'1i-clay _ AHi-aqueous + A1"15cav + AHSint [6] Summing just the fourth and fifth terms of equation 6 yields approximately the dehydration enthalpy (that is, the negative of the hydration enthalpy Athd) of the NAC, since hydration involves creating a cavity in water and placing the organic solute in the cavity (Goss and Schwarzenbach, 2001). Thus, AInlexpt : AH3cav + AI"13int + AHi-clay _ AI'lhyd + AHSint [7] 132 The purest measure for hydration of a compound is normally defined (Plyasunov and Shock, 2000) as transfer of the compound from its ideal gas (where it has no energetic interactions with its environment) to its aqueous solution. For example: 1,4-dinitrobenzene (g) <——> 1,4-dinitrobenzene (aq) [8] Note that this reaction is the sum of the dissolution reaction 1,4-dinitrobenzene (s) e 1,4-dinitrobenzene (aq) [9] and the reaction 1,4-dinitrobenzene (g) <—) 1,4-dinitrobenzene(s) [10] Reaction 10 is the reverse of the sublimation reaction, so the hydration enthalpy can be expressed as (Plyasunov and Shock, 2000) Athd = AHdis — AHsubr [11] Fortunately, estimates for sublimation enthalpies AH...“ have been published (Chickos and Acree, 2002) for several of the NACs studied here, and are listed in Table 3. This allows equation 7 to be rewritten AI'Iertpt = AH3cav + AH31nt + AHi-clay + AHsubl _ AHdis + AHSint [12] 133 Now, when Athd (equation 7) or AH...“ and AHd,-, (equation 12) can be looked up or estimated for any compound, it becomes easier to estimate an experimental adsorption enthalpy for that compound. Results and Discussion Algorithm 1 validation As a first test of our overall method and specific algorithms, we have computed adsorption enthalpies for the two nitrobenzene derivatives (R-NOz and R-COOCH3, where R=nitrobenzene) that adsorbed most strongly to the smectite in the experimental study (Boyd et al., 2001) and compared them with that of p-xylene, which is of similar size but lacks the polar functional groups that might enable complexation with interlayer cations so is expected to sorb with a much less favorable enthalpy (Table 1). 134 Avoom r? 6 So . -- 2 a at. own-8e- ode- 82% 35+ 0 H 8+ : a w? 82?-.. v H mm- om a w- omn-ooo- o.oo- own-oom o.oo+ mm a 53+ 2 H mm. mZu. - mm H o+ omn-ooo! o.oo- own-oom o.oo+ m: H 3 .+ E H on m:0OO0-m own- on: _Hw_- nmflm- 5H9»? E HVmT NOZ-m 3 ooo- o.oo- 8 oom+ o.oo+ 35w? as m< .8985 .938 8935 .033 oNcooobmcnmo scene-82 3383mm m :< m I4 m 34 m m< 3833-. :4- EE :4 2:89:00 65.388 :23 .8. can: 9.0.... m: E... «a... .5. 8.5838 2...; 25. .N ouawmm 6.... E «.93 as oouauwfioc Exams—co a... .3 1252. 21.2.5.5 .8339. we memo—«£5 69.3.3... 1.... 68.....th A «Bah. 135 Note that the clay-organic interaction enthalpies are quite favorable for all nitrobenzene solutes but not for p-xylene. The value of AH-wy for 1,4-dinitrobenzene adsorption to K- SWy-Z smectite was recently estimated to be -93 kJ/mol (Li et al., 2004), compared with our estimate (Table 1) of -1 54 kJ/mol. The same paper estimated AH-{iay for 1,3,5- trinitrobenzene adsorption to K-SWy-Z smectite at -124 kJ/mol (Li et al., 2004), again much smaller than our estimate (Table 1) of -235 kJ/mol. The only other estimates we know of for related systems are quantum mechanical: -38 kJ/mol for 1,3,5- trinitrobenzene interacting with a single, neutral 2:1 clay surface (Pelmenschikov and Leszczynski, 1999), and -49 kJ/mol for nitrobenzene interacting with two waters of hydration at the surface of a Na-smectite fragment (Gorb et al., 2000). The values of [AH-{lay for 1,4~dinitrobenzene and 1,3,5-trinitrobenzene computed here by molecular mechanics are about 65 and 90% larger, respectively, than the values estimated from experimental data (Li et al., 2004), and thus may indicate some deficiency in either the force field or the structural model used in our simulations. Electrostatic interactions account for >100% of AH-my for 1,4-dinitrobenzene (Table 2), so the forcefield perhaps overpredicts the tendency of 1,4-dinitrobenzene to form inner-sphere complexes with K: On the other hand, the molecular dynamics estimates for AH,.aqueous were done in the absence of any ionic species, yet the water-organic interaction enthalpies are also strong and favorable, so -AH-.,-queous values are large and positive for the nitrobenzenes. 136 Table 2. The F reundlich adsorption coefficients for substituted nitrobenzenes and the adsorption energies of the corresponding substituents. chemical Kf, Freundlich AHelectrostatic AHvdw AH-m AG of structure coefficient (l/g) solvation kJ/mol No,_©—NH, 0.0499 -137 40 -98 13.0 140,001+ 0.182 -91 5.0 -86 6.2 No,_©—oc-, 0.452 -75 17 -58 13.8 No,.©—cno 1-05 -107 6.0 -101 10.4 NOz—O—COOCH3 3.28 -103 -5.0 -108 14.4 rum—Gem”, 2-62 -104 5.0 -99 12.0 140,—ch 1.16 -129 24 -106 11.1 8102.0140- 2.87 -l67 14 -154 19.3 ”£001, 20 19 38 15.8 TNB -236 2.0 -235 15.8 The values of Aquucous for all solutes in Table 1 are within a few percent of their analogous values AHtc-ay in the clay phase. This is really quite surprising given that the clay interlayer contains a 12 mol/L solution of K+ while the bulk aqueous phase contained only water and the organic solute. From these results, we anticipate that Algorithm 1 will be difficult to make useful, because ABE-a, and AHi.aqueous differ by less than their standard deviations. Thus, the overall adsorption enthalpy AHexpt (a relatively small number) has to be predicted by subtracting two rather uncertain large numbers, a situation that does not allow for statistically significant predictive power. Literature estimates are available for comparison with some values of -AH.-aquc0u, in Table 1, but none of them are direct. Sandberg et al. (2002) predict 73 kJ/mol for -AG--aqucous of p-xylene. To convert this to a -AH--aque0us value, we note that Gallicchio et al. (2000) estimate that, for a range of 11 alkanes, the ratio of AHi.aqueous / AGi.aqueous averages 1.17 :t 0.04. This implies that -AH.-.aqucous should be near 85 kJ/mol for p-xylene. which is about twice our estimate. Estimates for the values of -AH-.aqueous for the substituted nitrobenzenes have to begin with nitrobenzene, since we know of no estimates for di- or trinitrobenzenes. The -AG,-.aqueous for nitrobenzene is estimated as 76 kJ/mol (Sandberg et al., 2002). Furthermore, adding a nitro-substituent to benzene, toluene, or phenol result in consistent changes of -AG-.;,qu¢0us of 23 kJ/mol (Sandberg et al., 2002), so reasonable guesses for -AG--aqueous values of 1,4-dinitrobenzene and 1,3,5-trinitrobenzene are 99 and 122 kJ/mol, respectively. Scaling these values by 1.17 (Gallicchio et al., 2000) converts them to crude approximations of 116 and 143 kJ/mol for the -AH,-aquc0u, of 1,4-dinitrobenzene and 1,3,5-trinitrobenzene, respectively, compared with 149 and 227 kJ/mol in Table 1. Our simulated estimates of -AH-.aqueous for p-xylene are too low while 138 those for the nitrobenzenes appear to be too large in comparison with previous estimates (Sandberg et al., 2002) that were also based on molecular simulations. A possible explanation is that force filed parameters used in either the previous study, the present study, or both were inaccurate or at least disagreed. For example, we used charges of +0.6170 on N and —0.4280 on O atoms in —N02 groups, while the previous study (Sandberg et al., 2002) used 0.58 and -0.33, respectively. Apparently, they (Sandberg et al., 2002) scaled the OPLS-AA forcefield (Price et al., 2001) charges (+0.65 and -0.37) by 0.9 in order to improve their fits to experimental free energies of hydration. Our previous quantum mechanical study (Boyd et al., 2001) of 15 small gas-phase nitroaromatic compounds fit atomic partial charges to the molecular electrostatic potentials and found that charges on N were +0.66 :t 0.01 and those on 0 were -O.4O i 0.01. In condensed phases like water, these charges would probably be larger in magnitude due to polarization caused by interactions with solvent molecules. Relevant estimates for the cavitation component of ABS have been published, but they are again indirect since AG5 values are more commonly available. Molecular dynamics simulations have been used several times, the first being (Postma et al., 1982), to estimate free energies for cavity formation in water, and the numbers come to a reasonable consensus. Taking cyclohexane as an example (density of the liquid at 20 °C is 0.779 g/cm3 (Lide, 1995) so molecular volume is 180 A3, similar to dinitrobenzene), three recent estimates for -AGS, the free energy required to form a cavity in water that can hold cyclohexane are 66 (Gallicchio et al., 2000), 64 (Sandberg et al., 2002), and 72 kJ/mol (Hdfinger and Zerbetto, 2003). Furthermore, Gallicchio et a1. (2000) estimated the enthalpic contribution for the first time and found that, for a range of eight C4- to C (,- 139 alkanes, the ratio of AH5 / AGs averages 0.58 i 0.03. Pertinent to Table 1, Sandberg et al. (2002) predict -68 kJ/mol for A65 of p-xylene, and using AH5 / A05 = 0.58 (Gallicchio et al., 2000) implies that the cavitation component of AH5 for p-xylene should be near -39 kJ/mol. Consider another way to make this estimate: The cavitation free energy A65 for a single water molecule is estimated at 15.6 i 0.2 kJ/mol (Sandberg et al., 2002). Converting this to AH5 using the 0.58 (Gallicchio et al., 2000) conversion factor yields 9.0 kJ/mol per water molecule. Given that the molecular volume of p-xylene (203 A3, (Lide, 1995) is about seven times that of water, and given that cavitation energies depend predominantly on molecular volume and surface area (Gallicchio et al., 2000; Hofinger and Zerbetto, 2003), the cavitation enthalpies AH5 for all the solutes considered here should be near -60 kJ/mol. A third way to estimate the cavitation energy is through use of an equivalent spherical radius. Again, cavitation energies like AH5 are thought to depend more on volume than on any particular shape (Hofinger and Zerbetto, 2003), and polynomials that are quadratic or cubic in equivalent spherical radius of the cavity often fit molecular estimates well (Hofinger and Zerbetto, 2003; Sandberg et al., 2002). For p- xylene, with a volume of 203 A3, the equivalent spherical radius is 3.65 A and a polynomial fit to alkanes (Sandberg et al., 2002) yields 100 kJ/mol for A65 to form that sphere in water. Applying AH5 / A65 = 0.58 (Gallicchio et al., 2000) again implies that AH5 for p-xylene should be about -58 kJ/mol. Since all the solutes of Table 1 are roughly the same size, -60 kJ/mol should be a reasonable estimate for all AH5 cavitation values and were also used for the cavitation component of AH5 in the clay since we know of no relevant estimates there. Thus we assume for now that cavitation energy differences in 140 the two phases will be insignificant, which looks like a reasonable assumption given the magnitudes of some of the other energy terms in Table 1. The second components (for the interaction of the cavity contents with its environment) of AH3 (the enthalpy required to take water out of a cavity in the clay interlayer) and ABS (the enthalpy change for filling an aqueous cavity with water — see Figure 2) require comment. These numbers were obtained by applying algorithm [5] to one or more water molecules, then normalizing the resultant interaction energy to the 10 water molecules considered to be leaving the interlayer cavity or entering the aqueous phase cavity. The range of values in Table 1 comes from ten or more estimates, each time choosing different subsets, for both the clay phase and the aqueous phase. The variability and overlap in these values that is evident in Table 1 would again seem to preclude their use as part of a sum to estimate relatively small adsorption energies. Are the estimates (Table l) for interaction of water inserted into an aqueous cavity reasonable? We know of no previous studies that tried to estimate AH3, but A115 values can be estimated. As discussed above, +9.0 kJ/mol is a reasonable estimate for the cavitation component of AH5, but what about this interaction component. The sum of the two components should just be the hydration enthalpy (Gallicchio et al., 2000), which is -44.0 kJ/mol (Lide, 1995). This implies by difference that the interaction component of AH5 for insertion of water into an aqueous cavity should be -53 kJ/mol. Our estimates of -60 to -72 kJ/mol are somewhat more negative than this value, but within reason. Our value may be a bit large because we did not allow the water molecules bordering the cavity to reorient their H-bonding patterns during our estimate: We simply deleted a cavity of water molecules and recalculated the energy. 141 The sum of AH3 and AH5 might be compared with experimental data, since AH3+AH5 should approximate the energy required to move water out of the smectite interlayer and into bulk water. That is, the sum of AH3+AH5 refers to the reaction (equation 2) C lay(aq) + cavitymm (—> Clay-cavity(aq) + 1120-3,” [13] which is equivalent to cavitym) <-—) cavity(c-ay) [14] which seems equivalent to H2()(clay) ('9 H20(aq) [1 5] The measured heat of immersion for the first layer of water loading on K- smectites is about -1 to -4 kJ/mol of water (Berend et al., 1995), for an average of about -2.5 kJ/mol, so AH; + AH5 should be about +2.5 kJ/mol of water. Thus, for the 10 water molecules used to represent each cavity, we would expect the magnitude of the insertion component of AH3 to be about 25 kJ/mol larger than that of AHs. Instead, there is little difference and, if anything, the central tendency of AH3 is to have a slightly smaller magnitude, although any differences are again swamped out by the overall variability. The preliminary results of Table 1 are encouraging in that a) our algorithm yields adsorption enthalpies of the correct sign for for agreement with the experimental data (Li et al., 2004) for 1,4-dinitrobenzene and 1,3,5-trinitrobenzene (and, for what it is worth, the standard deviations for the computed estimates overlap with the experimental 142 estimates), and b) adsorption enthalpies for nitroaromatics and p-xylene are predicted to be significantly different, even at this preliminary stage. A significant challenge in the path to making the algorithm USCfiJI is the size of the error bars at each step because they cause the overall Eexp- estimates to have an uncertainty that will ofien be larger than the value itself. Further work is needed to understand and minimize the sources of uncertainty in the energy estimates. The interaction energy was dominated by electrostatic rather than van der Waals terms (Table 2), in accordance with the earlier concepts that complexation interactions between interlayer cations, K+ in this case, and the nitroaromatic molecules are critical. These complexation energies also seem to dwarf the aqueous solvation energies of the nitroaromatics as a contributor to the overall adsorption energy. To test the validity of the algorithm 1, we used the data from Boyd et a1. (2001) (Figure 3). The numbers along the right hand side are our simulated interaction energies between organic and the smectite. Note that they follow a reasonable trend with strong interaction energy for strongly sorbing ones and smaller interaction energy for poorly sorbing ones, although some of them are out of order. But, in general, the trend of predicted energies for experimental sorption results is promising. 143 18 — 16 _ ~234zt17 g: 14 ~ a: ——TNB* 5.. 12 — ——p-N02 E 10 p-COOCH3 8 -163i16 p'ggfig3 8 8 - -108:t14 p' E 99:1:15 _p'OH is; 6 ‘ ' ——p-CN < 4 - —p-NH2 -106:l:18 2 K-58i16 " —-86:l:16 0 - 38:19 0 5 10 Equilibrium Conc., mglL Figure 3: Validity of algorithm with experimental data. For comparison, p-xylene, which is dimethylbenzene, which has an interaction energy of +38 kJ/mol. A positive number tells us that it is unfavorable and most of the nitro compounds are favorable. Xylene cannot form complexes and so its interaction energy is positive. (All the calculation energies are for monolayer systems where complexation interactions will be maximum.) All-m values were quite favorable for all nitroaromatic compounds but unfavorable for p—xylene. Thus, the algorithm shows promise for computation of useful AH-m values, but the adsorption energies depend on factors other than AH-m. For example, the energies for removal of the organic from aqueous solution must be a component of the adsorption energy, and this factor shows a range of 6 to 19 kJ/mol (see Table 2). In our simulations, the energies for removing organics from their cavities in bulk water were estimated at +149 i 21 kJ/mol for p-NOZ, +114 i 16 kJ/mol for p-COOC H 3, +227 1: 23 kJ/mol for TNB, and +40 3: 9 kJ/mol for p-xylene (Table 1). Therefore, adsorption enthalpies are estimated to be -5 i 27 for p-NOz, +6 i 22 for p-COOCH3, -8 i 29 for TNB, and +78 i 15 for p-xylene. A goal of each simulation was a 12.3 A d-spacing for the clay-water—organic system. Using p-COOCH3 as a test for the sensitivity of AHint, four different simulations were performed at three water contents that each resulted in a 12.3 A d-spacing (Figure 4). Variability in AHint was about 2 kJ/mol for a given water content but ~21 kJ/mol among the different water contents. Within a given simulation, standard deviations for AHim among the 150 observations were ~ 17 kJ/mol. 145 14.93 d-spacing, A 12.87 1225 12.31 1239 A Ar, c 15.31 12333 - ”-47 42 43 44 45 47 50 55 89 No. of interlayer waters Figure 4. Sensitivity Analysis. Algorithm 2 validation In applying algorithm 2, we can make use of our discovery using algorithm 1 that AHgm- and AH5-n- cancel each other (Table 1). This simplifies the estimation of an experimental adsorption enthalpy because only simulations of the clay phase are necessary, using AHexpt = AH3cav + AHi-clay + AI"Isubl - AI"Idis [16] Table 3 shows the results of applying algorithm 2 to the two nitroaromatic compounds for which we have measured the thermodynamics of adsorption to K-smectite (Li et al., 2004). Note that, while the predictions are not quantitative, they again have the proper sign and now have the advantage that the predicted adsorption enthalpy values themselves are larger than their standard deviations. 146 $8.0. :8 do so ._ A88 8284 use 8828 s : 4 mm 8+ 053-8 «How. 2 H 5. mm ~Hno~ S Hmmm- oo+ mZ-_. .3:- 2 4mm- 8 In; 2 32- 8+ Noz.Q axom< 3qu 988.802 8885-”— mar? “3324 8.8.34 23:4 8:53:80 anon—«5.8 8:988: 73.8.3898 «88:8 3 N Bantam—a .8 8=3=AE< .m 038,—. I47 Thus, we conclude that the method for simulating adsorption enthalpies shows promise. For example, predictions for nitroaromatic compounds versus p-xylene are significantly different. Further work is needed that will attempt to isolate and minimize sources of uncertainty to find out how precise the enthalpy predictions can become. 148 LITERATURE CITED Andersson, P.L., U. Maran, D. Fara, M. Karelson, and J.L.M. Hermens. 2002. General and class specific models for prediction of soil sorption using various physicochemical descriptors. Journal of Chemical Information and Computer Sciences 42:1450-1459. Atkins, P.W. 1998. Physical chemistry. 6th ed. W.H. Freeman and C 0., New York, NY. Bailey, G.W., and J .L. White. 1970. Factors influencing the adsorption, desorption, and movement of pesticides in soil. Residue Reviews 32:29-92. Berend, 1., J .M. Cases, M. Francois, J .P. Uriot, L. Michot, A. Masion, and F. Thomas. 1995. Mechanism of adsorption and desorption of water-vapor by homoionic montmorillonites .2. The Li+, Na+, K+, Rb+ and Cs+- exchanged forms. Clays and Clay Minerals 43:324-336. Bleam, W.F. 1993. Atomic theories of phyllosilicates: Quantum chemistry, statistical mechanics, electrostatic theory, and crystal chemistry. Reviews of Geophysics 31 :51-73. Boyd, S.A., G.Y. Sheng, B.J. Teppen, and CJ. Johnston. 2001. Mechanisms for the adsorption of substituted nitrobenzenes by smectite clays. Environmental Science & Technology 35:4227-4234. Briggs, G.G. 1981. Theoretical and Experimental Relationships between Soil Adsorption, Octanol-Water Partition-Coefficients, Water Solubilities, Bioconcentration Factors, and the Parachor. Journal of Agricultural and Food Chemistry 29:1050- 1059. Chang, F .R.C., N.T. Skipper, and G. Sposito. 1997. Monte Carlo and molecular dynamics simulations of interfacial structure in lithium-montmorillonite hydrates. Langmuir 13:2074-2082. Chickos, J .S., and WE. Acree. 2002. Enthalpies of Sublimation of Organic and Organomettalic Compounds. 1910-2001. Journal of Physical and Chemical Reference Data 31:537-698. Chiou, CT. 1998. Adsorption and partition of organic pollutants and pesticides in soils Encyclopedia of Environmental Analysis and remediation. John Wiley and Sons, New York, NY. Chiou, C.T., L.J. Peters, and V.H. Freed. 1979. Physical concept of soil—water equilibria for nonionic organic compounds. Science 206:831-832. 149 Chiou, C.T., P.E. Porter, and D.W. Schmedding. 1983. Partition Equilibria of Non-Ionic Organic-Compounds Between Soil Organic-Matter and Water. Environmental Science & Technology 17:227-231. Cygan, RT 1998. Molecular models of metal sorption on clay minerals, p. 127-176, In J. D. Kubicki and W. F. Bleam, eds. Molecular modeling of clays and mineral surfaces. Clay Minerals Society, Boulder, CO. Cygan, RT. 2001. Molecular modeling in mineralogy and geochemistry. Reviews in mineralogy and geochemistry 42:1-35. Delville, A. 1991. Modeling the Clay Water Interface. Langmuir 72547-555. Gallicchio, 13., M.M. Kubo, and RM. Levy. 2000. Enthalpy—entropy and cavity decomposition of alkane hydration free energies: numerical results and implications for theories of hydrophobic solvation. Journal of Physical Chemistry B 104:6271-6285. Gawlik, B.M., N. Sotiriou, E.A. Feicht, S. SchulteHostede, and A. Kettrup. 1997. Alternatives for the determination of the soil adsorption coefficient, K-oc, of non- ionicorganic compounds - A review. Chemosphere 34:2525-2551. Gerstl, Z. 1990. Estimation of organic chemical sorption by soils. Journal of Contaminant Hydrology 6:357-375. Gerstl, Z., and U. Mingelgrin. 1984. Sorption of Organic-Substances By Soils and Sediments. Journal of Environmental Science and Health Part B-Pesticides Food Contaminants and Agricultural Wastes 19:297-312. Gohlke, H., and G. Klebe. 2002. Approaches to the description and prediction of the binding affinity of small-molecule ligands to macromolecular receptors. Angewandte Chemie-Intemational Edition 41 :2645-2676. Gorb, L., J .D. Gu, D. Leszczynska, and J. Leszczynski. 2000. The interaction of nitrobenzene with the hydrate basal surface of montmorillonite: an ab initio study. Physical Chemistry Chemical Physics 2:5007-5012. Goss, K.U., and RP Schwarzenbach. 2001. Linear free energy relationships used to evaluate equilibrium partitioning of organic compounds. Environmental Science & Technology 3521-9. Gramatica, P., M. Corradi, and V. Consonni. 2000. Modelling and prediction of soil sorption coefficients of non-ionic organic pesticides by molecular descriptors. Chemosphere 41 :763-777. Greathouse, J .A., K. Refson, and G. Sposito. 2000. Molecular dynamics simulation of water mobility in magnesium- smectite hydrates. Journal of the American Chemical Society 122:11459-1 1464. 150 Guo, R.B., X.M. Liang, J .P. Chen, C. Zhang, and A. Kettrup. 2002. Prediction of soil adsorption coefficients from retention parameters on three reversed-phase liquid chromatographic columns. Analytical Chemistry 74:655-660. Haderlein, SB, and RP. Schwarzenbach. 1993. Adsorption of substituted nitrobenzenes and nitrophenols to mineral surfaces. Enviromnental Science & Technology 27:316-326. Haderlein, S.B., K.W. Weissmahr, and RP. Schwarzenbach. 1996. Specific adsorption of nitroaromatic explosives and pesticides to clay minerals. Environmental Science & Technology 30:612-622. Halperin, 1., BY. Ma, H. Wolfson, and R. Nussinov. 2002. Principles of docking: An overview of search algorithms and a guide to scoring functions. Proteins-Structure Function and Genetics 47 :409-443. Hdfinger, S., and F. Zerbetto. 2003. On the cavitation energy of water. Chemistry - A European Journal 9:566-569. Huuskonen, J. 2003. Prediction of soil sorption coefficient of a diverse set of organic chemicals from molecular structure. Journal of Chemical Information and Computer Sciences 43: 1457-1462. J aynes, W.F., and SA. Boyd. 1991. Hydrophobicity of siloxane surfaces in smectites as revealed by aromatic hydrocarbon adsorption from water. Clays and Clay Minerals 39:428-436. Johnston, C.T., M.F. De Oliveira, B.J. Teppen, G.Y. Sheng, and SA. Boyd. 2001. Spectroscopic study of nitroaromatic-smectite sorption mechanisms. Environmental Science & Technology 35:4767-4772. Johnston, C.T., G. Sheng, B.J. Teppen, S.A. Boyd, and M.F. De Oliveira. 2002. Spectroscopic study of dinitrophenol herbicide sorption on smectite. Environmental Science & Technology 36:5067-5074. Kalinichev, A.G., R.J. Kirkpatrick, and RT. Cygan. 2000. Molecular modeling of the structure and dynamics of the interlayer and surface species of mixed-metal layered hydroxides: Chloride and water in hydrocalumite (Friedel's salt). American Mineralogist 85: 1046-1052. Kanazawa, J. 1989. Relationship between the Soil Sorption Constants for Pesticides and Their Physicochemical Properties. Environmental Toxicology and Chemistry 81477-484. Karickhoff, SW. 1981. Semiempirical Estimation of Sorption of Hydrophobic Pollutants On Natural Sediments and Soils. Chemosphere 10:833-846. 151 Karickhoff, S.W., and KR. Morris. 1985. Sorption Dynamics of Hydrophobic Pollutants in Sediment Suspensions. Environmental Toxicology and Chemistry 42469-479. Karickhoff, S.W., D.S. Brown, and TA. Scott. 1979. Sorption of Hydrophobic Pollutants On Natural Sediments. Water Research 13:241-248. Laird, DA. 1997. Bonding between polyacrylamide and clay mineral surfaces. Soil Science 162:826-832. Lasaga, AC, and G.V. Gibbs. 1990. Ab initio quantum-mechanical calculations of surface reactions - A new era?, p. 259-289, In W. Stumm, ed. Aquatic chemical kinetics: Reaction rates of processes in natural waters. John Wiley and Sons, Inc., New York, NY. Li, H., B.J. Teppen, C.T. Johnston, and SA. Boyd. 2004. Thermodynamics of nitroaromatic compound adsorption from water by smectite clay. Environmental Science & Technology 38:5433-5442. Lide, DR. 1995. Handbook of Chemistry and Physics. 76th ed. CRC Press, Boca Raton, FL. Lohninger, H. 1994. Estimation of Soil Partition-Coefficients of Pesticides from Their Chemical-Structure. Chemosphere 29: 161 1-1626. Maple, J .R., M.-J. Hwang, T.P. Stockfish, U. Dinur, M. Waldman, C.S. Ewig, and A.T. Hagler. 1994. Derivation of Class H force fields. 1. Methodology and quantum force field for the alkyl functional group and alkane molecules. Journal of Computational Chemistry 1 5: 162-1 82. Accelrys, Inc. 2000. Cerius2 Molecular Simulation Software, Release 4.2. Accelrys, Inc., San Diego, CA. Muller, M., and W. Kordel. 1996. Comparison of screening methods for the estimation of adsorption coefficients on soil. Chemosphere 32:2493-2504. Pelmenschikov, A., and J. Leszczynski. 1999. Adsorption of 1,3,5-trinitrobenzene on the siloxane sites of clay minerals: Ab initio calculations of molecular models. Journal of Physical Chemistry B 103:6886-6890. Plyasunov, A.V., and EL. Shock. 2000. Thermodynamic functions of hydration of hydrocarbons at 298.15 K and 0.1 MPa. Geochimica Et Cosmochimica Acta 64:439-468. Postrna, J .P.M., H.J.C. Berendsen, and J .R. Haak. 1982. Thermodynamics of cavity formation in water. A molecular dynamics study. Faraday Symposia of the Chemical Society 17:55-67. 152 Price, M.L.P., D. Ostrovsky, and W.L. Jorgensen. 2001. Gas-Phase and Liquid-State Properties of Esters, Nitriles, and Nitro Compounds with the OPLS-AA Force Field. Journal of Computational Chemistry 22: 1340—1352. Sabljic, A., H. Gusten, H. Verhaar, and J. Hermens. 1995. Qsar Modeling of Soil Sorption - Improvements and Systematics of Log K-Oc Vs Log K-Ow Correlations. Chemosphere 31 :4489-4514. Sandberg, L., R. Casemyr, and O. Edholm. 2002. Calculated hydration free energies of small organic molecules using a nonlinear dielectric continuum model. Journal of Physical Chemistry B 106:7889-7897. Sauer, J ., and R. Ahlrichs. 1990. Gas phase acidities and molecular geometries of H3SiOH, H3COH, and H20. The Journal of Chemical Physics 93:2575-25 83. Schafer, L., S.Q. Kulpnewton, K. Siam, V.J. Klimkowski, and C. Vanalsenoy. 1990. Abinitio Studies of Structural Features Not Easily Amenable to Experiment. Theochem-Joumal of Molecular Structure 68:373-385. Schmidt, T.C., P. Kleinert, C. Stengel, K.U. Goss, and SB. Haderlein. 2002. Polar fuel constituents: Compound identification and equilibrium partitioning between nonaqueous phase liquids and water. Environmental Science & Technology 36:4074—4080. Sheng, G.Y., C.T. Johnston, B.J. Teppen, and SA. Boyd. 2001. Potential contributions of smectite clays and organic matter to pesticide retention in soils. Journal of Agricultural and Food Chemistry 49:2899-2907. Sheng, G.Y., C.T. Johnston, B.J. Teppen, and SA. Boyd. 2002. Adsorption of dinitrophenol herbicides from water by montmorillonites. Clays and Clay Minerals 50:25-34. Skipper, NT. 1998. Computer simulation of aqueous pore fluids in 2 : 1 clay minerals. Mineralogical Magazine 62:657-667. Skipper, N.T., K. Refson, and J .D.C. McConnell. 1991. Computer simulation of interlayer water in 2:1 clays. The Journal of Chemical Physics 94:7434-7445. Smith, DE. 1998. Molecular computer simulations of the swelling properties and interlayer structure of cesium montmorillonite. Langmuir 14:5959-5967. Sun, H. 1998. COMPASS: An ab initio force-field optimized for condensed-phase applications - Overview with details on alkane and benzene compounds. Journal of Physical Chemistry B 102:7338-7364. Sutton, R., and G. Sposito. 2001. Molecular simulation of interlayer structure and dynamics in 12.4 angstrom Cs-smectite hydrates. Journal of Colloid and Interface Science 237: 174-184. 153 Tao, 8., HS. Piao, R. Dawson, X.X. Lu, and H.Y. Hu. 1999. Estimation of organic carbon normalized sorption coefficient (K-oc) for soils using the fragment constant method. Environmental Science & Technology 33:2719-2725. Teppen, B.J., C.H. Yu, D.M. Miller, and L. Schafer. 1998. Molecular dynamics simulations of sorption of organic compounds at the clay mineral/aqueous solution interface. Journal of Computational Chemistry 19:144-153. Teppen, B.J., K. Rasmussen, P.M. Bertsch, D.M. Miller, and L. Schafer. 1997. Molecular dynamics modeling of clay minerals .1. Gibbsite, kaolinite, pyrophyllite, and beidellite. Journal of Physical Chemistry B 101:1579-1587. Weissmahr, K.W., M. Hildenbrand, R.P. Schwarzenbach, and SB. Haderlein. 1999. Laboratory and field scale evaluation of geochemical controls on groundwater transport of nitroaromatic ammunition residues. Environmental Science & Technology 33:2593-2600. Weissmahr, K.W., S.B. Haderlein, R.P. Schwarzenbach, R. Hany, and R. Nuesch. 1997. In situ spectroscopic investigations of adsorption mechanisms of nitroaromatic compounds at clay minerals. Environmental Science & Technology 31:240-247. Zhu, D., B.E. Herbert, M.A. Schlautman, E.R. Carraway, and J. Hur. 2004. Cation-pi bonding: A new perspective on the sorption of polycyclic aromatic hydrocarbons to mineral surfaces. Journal of Environmental Quality 33:1322-1330. 154 1111111111111111111