You are on page 1of 22
CIRCULATION COP ” ee un. UCRL 85582 i PREPRINT CHEMICAL MODELING OF IRREVERSIBLE REACTIONS IN NUCLEAR WASTE-WATER-ROCK SYSTEMS Thomas J. Wolery THIS PAPER WAS PREPARED FOR SUIIBITTAL TO Waste/Rock Interaction Technical Workshop Battelle Pacific Northwest Laboratories Seattle, Washington February, 1981 ; This is a preprint of @ paper intended for publicatio 4 ina journal of proceedings. Since changes may be nad before publication, this preprint is made available with the understanding that it will not be cited or reproduced without the permission of the author. Unclassified DISCLAIMER This document was prepared as an account of work sponsored by an agency of the United States Government. Neither the United States Government nor the University of California nor any of their employees, makes any warranty, express or implied, or assumes any legal liability or responsibility for the accuracy, completeness, or usefulness of any information, apparatus, product, or process disclosed, or represents that its use would not infringe privately owned rights. Reference herein to any specific commercial products, process, or service by trade name, trademark, manufacturer, or otherwise, does not necessarily constitute or imply its endorsement recommendation, or favoring of the United States Government or the University of California. The views and opinions of authors expressed herein do not necessarily state or reflect those of the United States Government or the University of California, and shall not be used for advertising or product endorsement purposes. CHEMICAL MODELING OF IRREVERSIBLE REACTIONS IN NUCLEAR WASTE-WATER-ROCK SYSTEMS ‘Thomas J. Wolery Lawrence Livermore National Laboratory Earth Sciences Division L-202, P.O. Box 808 Livermore, CA. 94550 : ABSTRACT Chemical models of aqueous geochemical systems are usually built on the concept of thermodynamic equilibrium. Though many elementary reactions in a geochemical system may be close to equilibrium, others may not be. Chemical models of aqueous fluids should take into account that many aqueous redox reactions are among the latter. The behavior of redox reactions may critically affect migration of certain radionuclides, especially the actinides. In addition, the progress of reaction in geochemical systems requires thermodynamic driving forces associated with elementary reactions not at equilibrium, which are termed irreversible reactions. Both static chemical wodels of fluids and dynamic models of reacting systems have been applied to a wide spectrum of problems in water-rock interactions. Potential applications in nuclear waste disposal range from problems in geochemical aspects of site evaluation to those of waste-vater-rock interactions. However, much further work in the laboratory and the field will be required to develop and verify such applications of chemical nodeling. INTRODUCTION Modeling is a means of developing and demonstrating understanding. "Chemical modeling" as used in the field of geochemistry generally refers to mathematical modeling of water-rock interactions. There is a broad spectrum of approaches and phenomena that fall under this banner (see the ACS Symposium Volume edited by Jenne 1979). Many of these approaches have potential application in nuclear waste disposal. One such application is the evaluation of Water-rock interactions in natural, unperturbed groundwater systems. This is a necessary part of repository site evaluation. Similar applications may be developed to predict the geochemical behavior of systems about an unbreached repository. Last, chemical modeling capabilities can be developed to study waste-water-rock systems to aid in predicting the consequences of repository breaching. Chemical modeling in geochemistry is largely aimed at on applying equilibrium thermodynamics to determining the state of aqueous fluids. These fluids include river and lake waters, sea water, ground waters and hydrothermal fluids. These models generally attempt to answer two questions: (1) how are the elementary solutes (such as Na*, Ca”*, HCO}, etc.) distributed among free ions, ion pairs, and complexes and (2) is the fluid saturated, undersaturated, or super-saturated with respect to various minerals? Chemical models of this type are generated by applying computer programs such as WATEQ, SOMLNEQ, EQ3, REDEQL, MINEQL, and others. Nordstrom et al. (1979a) discuss in more detail this type of chemical modeling and review most of the computer codes used for this purpose in geochemistry. Chemical models that are based entirely on the assumption of thermodynamic equilibrium usually compare ell in an overall sense with actual systems, yet often show signnificant discrepancy on specific points. The assumption of thermodynamic equilibrium is appropriate for reactions that adjust quickly to any perturbations. Ion pairing and most complexing reactions among aqueous solutes fall into this cla However, the equilibrium assumption is not alvays appropriate for oxidation-reduction reactions (even those involving only components of the aqueous phase), heterogeneous reactions (such as mineral dissolution and precipitation) reactions, and perhaps (depending on the time scale of interest) polymerization reactions. ‘This paper has two purposes. The first is to briefly review some aspects of chemical modeling which may have significant impact on programs for nuclear waste disposal. The second is to point out some of the recent work performed at our Laboratory under the auspices of the Waste-Rock Interactions Technology (WRIT) Program. AQUEOUS REDOX DISEQUILIBRIA VS. EH The aqueous oxidation-reduction (redox) reactions offer several well-documented examples of disequilibrium that is both pronounced and persistent (on the scale of hours to years). One simple example is the coexistence of dissolved oxygen and dissolved organic compounds in many natural surface waters at levels that far exceed thermodynamic equilibrium. The approach to redox equilibrium at earth surface temperature is often governed by biological processes due to the Slowmess of inorganic mechanisms. For example, dissolved sulfate commonly reduced biologically in marine sediments (see Berner 1971, Ch. 7). Yet inorganic reduction by as strong a reducing agent as hydrogen gas is not detectable at temperatures much less than 250 C (Malinin and Khitarov 1969). Redox reactions may be considered as the sum of half-reactions or redox couples (see, for example, Garrels and Christ 1965, ch. 1). If all redox couples in aqueous solutions were at equilibrium the theoretical Eh or redox potential of each couple would be the same. One might then hope to measure this with a platinua electrode and use the result as part of the input to a chemical model in order to distribute the total concentration of a multivalent solute (such as dissolved iron) over its various valence forms (ferrous and ferric in this example). Most of the computer codes reviewed by Nordstrom et al. (1979a) can do this. This approach often fails by yielding unrealistic estimates of aqueous species distribution for multivalent solutes and consequently incorrect estimates of saturation states for minerals composed of these components. One reason this approach can fail is that the platinum electrode may fail to respond properly when immersed in the solution (see Nordstrom et al. 1979b). A second resond is more fundamental: such use of the measured Eh parameter assumes that all the redox couples in solution are in equilibrium with each other. The examples of disequilibrium noted above demonstrate that this assumption can be faulty. A much more realistic practice is to use chemical modeling codes to estimate the theoretical redox potential of each major couple. This requires measuring the concentration or activity of the major multivalent solutes in each valence state; e.g., separate analyses for ferrous and ferric iroa, not just total iron. One can then determine whether or not the various redox couples are responding to one another (and also the platinum electrode). The utility of this approach has been demonstrated by Thorstenson (1970), Nordstrom et al. (1979b), and others. Several of the distribution-of-species codes, including WATEQ and Q3, have options to treat redox reactions in this manner. However, no such capability yet exists in any reaction-path code. MODELING EVOLUTION OF AQUEOUS WATER-ROCK AND WASTE-WATER-ROCK SYSTEMS Reaction-path models attempt to predict the irreversible evolution of a reacting aqueous geochemical system. An irreversible geochemical process can be thought of as representing the combined progress of @ set of linearly independent elementary chemical reactions. Some of these are not at equilibrium (are irreversible) and provide the thermodynamic driving force for the overall process. Other reactions may proceed reversibly; they adjust to maintain themselves in a state of thermodynamic equilibrium as the irreversible elementary reactions progress and perturb the state of the system. This is the concept of partial equilibrium. Helgeson (1968) used this concept to develop a general model of irreversible reactions in rock-water systems. This model was developed in the form of a computer program called PATHI (see also Helgeson et al. 1970). It was applied to such geochemical processes as diagenesis and metasomatism (Helsegon et al- 1969) and the genesis of ore-bearing fluids (Helgeson 1970). Helgeson's model assumed that the dissolution of the primary mineral reactants was rate-limiting (i.e., irreversible), and that the remaining reactions, including the aqueous redox reactions and the precipitation and dissolution of secondary minerals, were reversible Ges, each in a state of thermodynamic equilibrium at any point of the overall process). The relationship of this model to the tine variable was entirely arbitrary. The progress of the overall process was measured instead by the extent to which the primary reactants had been consumed. The systems that could be treated included both closed systems and open flow-through systems. Wolery (1978, 1979a) wrote a new chemical modeling software package composed of a distribution-of-species code (EQ3), an irreversible reaction-path code (EQ6), and a set of supporting thermodynamic data files (largely inherited from Helgeson's earlier work). The differences between EQ3 and other such codes vere discussed by Wolery (1979a) and Nordstrom et al., (1979a). £Q3 is required to initialize £96; it provides a chemical model of the state of an aqueous fluid. Wolery (1979a) discussed the improvements of EQ6 over the earlier PATH. Q6 was originally written to model reaction of sea water with basalt in hydrothermal systems at mid-ocean ridges (Wolery 1978, 1979). It has also been applied to modeling mineral precipitation in reinjection of spent geothermal fluids (Taylor et al. 1978) and formation of copper ore bodies (Brishall 1980). Wolery (1980) used the EQ3 and EQ6 codes to calculate models of the dissolution of idealized uraninite (U0,(,)) in oxidizing, noderately oxidizing, and reduced plutonic (granitic) ground waters at temperatures from 25 C to 200C. ‘These models assumed redox equilibrium in aqueous solution, which is in general a questionable assumption. In these systems the only significant redox couples were dissolved oxygen-vater ana u'*-U05*, which are unlikely to be grossly out of equilibrium with each other. Sulfate was present, but reduction to sulfide was quantitatively insignificant even assuming redox equilibrium. The thermodynamic data employed for uranium species were those for 25 © reported by Langmuir (1978), extrapolated to high temperature according to algorithms of Criss and Cobble (1964a, 1964b) and Helgeson (1967) by the computer codes of Barner and Scheuerman (1978). The accuracy of these algorithas breaks down as temperature increases. Consequently, model calculations based on their results for temperatures greater than 200 C should be considered speculative. In Wolery's (1980) simulations the speciation of dissolved uranium was always dominated by uranyl carbonate complexes up to 200 C. Simulations for 300 ¢ predicted that uranyl sulfate complexes would predominate. The 300 ¢ results suggest the need for new experimental measurenents. In general, though, current extrapolation techniques should be sufficient for most waste disposal needs. ‘The major needs with respect to thermodynamic data for chemical modeling in the context of nuclear waste isolation are updating the set of existing data and adding new data for elements not in the data set. ‘The major vaste component elements (actinides, rare earths, cesium and strontium) are poorly represented in most thermodynamic data sets, as are several other elements which are potential waste form components, such as titanium and zirconium, (components of SYNROC; Ringvood 1978). The distribution-of-species code WATEQ2 (Ball et al. 1980) has one of the better geochemical data bases now available. Its data base is well-documented and also fairly comprehensive for water-rock (but not waste-water-rock) systems. The members of the present generation of geochemical modeling codes are largely restricted to treating minerals with fixed compositions. Some codes (e-g., £Q3 and £Q6) are programmed to treat a few simple models of solid solution, but in practice these are not very useful due to the general lack of the additional required thermodynamic data. Some of the elements of major concern in nuclear waste isolation, e-g-, strontium and cesium, largely occur in natural systems ‘as dilute guest components in solid solution minerals (such as feldspars and clays). Helgeson and Aagaard (1970) proposed bringing the element of time into codes such as EQ6 by incorporating kinetic rate expressions based on transition-state theory (Aagaard and Helgeson 1977). Wolery (1980) discussed the consequences of such an approach to nuclear waste disposal studies. These include the possibility of long-term predictions of waste-water-rock interactions. RECENT WORK ON THE Q3/EQ6 CODES The £Q3/EQ6 chemical modeling codes are now being developed for application to nuclear waste disposal problems under the auspices of the Waste-Rock Interactions Technology (WRIT) Program. Two tasks were begun in pid-FYBO. The first was some code development of EQ6 to eliminate occasional abnormal termination of program execution when the redox variable (oxygen fugacity, Eh, or pe”) was initially unstable (poorly poised). The second was to begin updating and expanding the thermodynamic data base that supports EQ3 and EQ6 calculations. REDOX CODE DEVELOPMENT Atteapts to model reaction-paths that began with non-oxidizing ground waters frequently terminated abnormally at the beginning of the run. Non-oxidizing ground waters typically contain low concentrations of poising solutes (examples of which are dissolved oxygen and aqueous sulfide). In nature, water-rock interactions may provide most of the poising for such ground waters. This problem was solved by two actions. First, the arithmetic in £Q3 and EQ6 vas revised to minimize round-off/teuncation error. The redox variable (oxygen fugacity, Eh, or pe) is an input parameter to £93, but it is iteratively recomputed in EQ6. The value carried over from £Q3 is used only as a trial value. Round-off/truncation errors in each code were found to exert deleterious effects on the recomputation of the redox variable in the case of poorly-poised ground waters. A partial sort-suaming technique was applied to lengthy summations in order to eliminate this problem with minimal loss of efficiency. The other action was to create a capability in £Q6 to scan trial values of the redox variable over the range of stability of liquid vater. This feature originally showed the need for the more accurate arithmetic discussed above. Without the partial sort-summing, EQ6 was able to converge to a somewhat different redox value than had been input to £Q3 for several cases of ill-poised ground waters. The scanning technique is now also the method that gets EQ6 simulations over large jumps in the redox parameter that occur when a poising component (often dissolved oxygen) is exhausted. DATA BASE EXPANSION AND REVISTON The elements uranium, plutonium, strontium, cesium, fluorine, and Phosphorus were the focus of recent work to expand and revise the £Q3/£Q6 data base. Most of the data collected from the literature pertain to 25 C and require extrapolation to higher temperatures. Standard methods for doing this (Criss and Cobble 1964a, 1964b; Helgeson 1969; Helgeson et al. 1978) are adequate to about 200 C. Most of the existing data for conmon rock-forming minerals in the Q3/EQ6 data base are generated from the SUPCRT data base (Helgeson et al. 1978, and references therein). ‘These have been updated to be consistent with recent SUPCRT changes (H. C. Helgeson, written communication, 1980). Wolery (1980) put uranium in the data base, adopting the critical compilation of Langmuir (1978). Lemire and Tremaine (1980) later pointed out that Langmir did not include the aqueous complexes vo, (OW) and (UO,),(OH);; these were added to the data base. We retained Langmuir's data in those few cases where Lemire and Trenaine disagree with Langmuir. Plutonium was recently put in the £Q3/EQ6 data base by adopting the critical compilation of Lemire and Tremaine (1980). Limited data for strontium and cesium have been added to the data base. The geochemistries of these elements are strongly influenced by their roles as dilute guest components of major rock-forming minerals. Code development and thermodynamic data for mineral solid-solution is a necessary prerequisite to adequately model the behavior of these elenents, Fluorine and phosphorus are important elements to include in geochemical data bases because fluoride and phosphate significantly complex many cations. Ball et al. (1980) summarize many data for aqueous and mineral species of these elements. Their work has been used as a springboard to develop revised and extended data sets for the £Q3/£Q6 data base. Our work on these elements will be reported in detail elsewhere. EXAMPLE: URANIUM IN A HANFORD GROUNDWATER The use of the EQ3 code is illustrated by predicting the speciation of uranium and fluorine in a high pH-high fluoride ground water from the Hanford basalt. The ground water composition used here (Table 1) is based on a suite of analyses (not all of which are consistent) for well DC-2 reported by Apps et al. (1979). This calculation follows the revision of uranium data and addition of fluorine data discussed above. TABLE 1. Composition of a Hanford Ground Water (after Apps et al., 1979) Solute Cones) pps na* 18 i 2 ca* 0.26 sio, 153 a 95 ee a. 095 % oo§ a7 30, 26 v 0.001 pH 9.8 En 46 av Temp. 36 The speciation of dissolved uranium (Table 2) is dominated by uranyl carbonate complexing. This is to be expected in most ground waters (Langmir 1978). The UO,(OH)? aqueous complex is more important than most other uranium species, but is minor in comparison with the uranyl di- and tri-carbonate species. TABLE 2. Partial List of Aqueous Uranium and Fluoride Species species Cones, Molal U (total) 4.20 x 109 F (total) att x 1079 ut 3.45 x 10749 vt 9.59 x 1077 09" 1.81 x 10720 v0, 4,83 x 10717 v0, (c0,)° 5.98 x 10°!4 2 cone” 10 v0,,(605)7 2.88 x 10 v0, (C0,)5_ 3.91 x 107° ao 12 v0, (0H)? 1.03 x 10 uCoH. 3.67 x 103 6 18 OF 1.67 x 10 2° 17 0,89 1.06 x 10 v0,F, 2.82 x 10717 vo,F 6.35 x 10720 Fy 7 r 1.10 x 10° ne? 2.86 x 1071 HFS 1.05 x 1071? 2, 19 HF 1.46 x 10 23 6 NaF 1.21 x 10 cart 5.29 x 1078 Fluoride is precluded from being an effective complexer of uranium in most ground waters because it is largely tied up by hydrogen ion. In this high-pH ground water, however, fluoride occurs mostly as free anion. It is still ineffective in complexing uranium. Uranous fluoride complexes, not included in Table 2, all exist many orders of magnitude below the molecule per liter level. What would happen to this ground water if UO,’_) (uraninite) 2¢c! were to irreversibly dissolve in it? The EQ6 code can be used to scope out the consequences. Fig. 1 depicts changes in Eh, pH, and total dissolved uranium as a function of the amount of UO,,. that has 2c) dissolved. It is not possible at the present time to predict these changes as a function of time (see Wolery 1980). Eh drops progressively, pl remains nearly constant, and total dissolved uranium increases until the solution becomes UO, (c) saturated and reaction stops. No secondary mineral products form in this simlation. In similar simulations of U0, ,) dissolution in oxidizing granitic ground waters (Wolery 1980) schoepite (v0,(OH)y H,0) or beta-U0,(OH),, alpha-U,0,, and U,0,(c) appeared as secondary products, all transient except U,0,(c). Changes in dissolved uranium speciation also occur as U0,¢<) dissolution progresses (Fig. 2). The most dramatic feature is the increase in the concentration of U(OH);, which becomes the most abundant uranium species near the end of the reaction. 7 initiol Eh _ ~ ~-l00 U02(e) = saturates Pr -200 — 103) x o 95] =_ “15 + > -80 € g g -as 40 a “3.7 UO, dissolved, log mol/kg HO FIGURE 1. Changes in Eh, pH and Total Dissolved Uranium During Dissolution of UO, ae) it Hanford Ground Water. ae T T a 5 (CO3) 3° U02 (605) (C03) £ +0 | 2 = U(OH)S UO2 {c) -uZ saturates | “2 = 1 1 a “10 “8 “7 UOg dissolved, log mol/kg H20 FIGURE 2. Speciation of Dissolved Uranium During Dissolution of VO, ,_) in Hanford Ground Water 2¢e) ‘SUMMARY Chemical models of aqueous geochemical systems are useful tools for studies related to nuclear waste disposal in geologic media. Such studies include those of rock-water systems (geochemical aspects of repository site evaluation and consequences of unbreached repositories) and waste-water-rock systems (consequences of breached repositories). Most chemical models depend considerably on assumptions of thermodynamic equilibria. These assumptions should be carefully examined for their appropriateness, especially in the case of oxidation-reduction reactions. The usefulness of chemical modeling to the nuclear waste disposal effort and to aqueous geochemistry in general is limited by the scope and quality of both thermodynamic and kinetic data bases. Increasing the degree to which chemical models mirror reality requires expanding and refining these data bases and reducing the number of inappropriate assumptions of thermodynamic equilibrium in favor of kinetic constraints. Very few studies (such as Aagaard and Helgeson, 1977) have been done thus far to develop general rate law expressions to describe the kinetics of the irreversible reactions that are important in ground water systems. ‘There is a great need for laboratory studies of the dissoluticn kinetics of the minerals that make up rocks at potential repository sites and of the kinetics of aqueous redox reactions under conditions found in ground water systems. Code development, experimentation, and field studies are all important parts of developing and testing chemical models. ACKNOWLEDGEMENTS I thank David Coles, Dana Isherwood, and Kevin Knauss for their comments on this paper. REFERENCES Aagaard, P., and Helgeson, H. C. 1977. "Thermodynamic and Kinetic Abstract, Geol. Constraints on the Dissoltuion of Feldspars Soc. Amer. Abstracts with Programs, 9, 873. Apps, J-, et al. 1979. Geohydrological Studies for Nuclear Waste Isolation at the Hanford Reservation, Volumes 1 and 2. LBL~8764, Lawrence Berkeley Laboratory, Berkeley, California. Ball, J. W., Nordstrom, D. K., and Jenne, E, A. 1980. Additional and Revised Thermomechanical Data and Computer Code for WATEQ? - - A Computerized Chemical Model for Trace and Major Element Speciation ‘and Mineral Equilibria of Natural Waters. U.S. Geological Survey Water Resources Investigations, 78-116. Barner, H. E., and Scheuerman, R. V. 1978. Handbook of Thermochemical Data for Compounds and Aqueous Species. John Wiley and Sons, New York. Berner, R. As 1971. Principles of Chemical Sedimentology. McGraw-Hill Book Company, New York. Brinhall, G. H., Jr. 1979. "Deep Hypogene Oxidation Porphyry Copper Potassiua-Silicate Protore at Butte, Montana: A Theoretical Evaluation of the Copper Remobilization Hypothesis". Econ. Geol. 75: 384-409. Criss, C. Me, and Cobble, J. W. 1964a. “The Thermodynamic Properties of High Temperature Aqueous Solutions. IV. Entropies of the Ions up to 200 C and the Correspondence Principle." J. Amer. Chem. Soc. 86: 5385-5390. Criss, c. M., and Cobble, J. W. 1964. "The Thermodynamic Properties of High Temperature Aqueous Solutions. V. The Calculation of Tonic Heat Capacities up to 200 C. Entropies and Heat Capacities Above 200 6." J. Azer. Chem. Soc., 86: 5390-5393. Garrels, R. M., and Christ, C. L. 1965. Solutions, Minerals and Equilibria. Freeman, Cooper, and Company, San Francisco. Helgeson, H. C. 1967. "Thermodynamics of Complex Dissociation in Je Phys. Chem. 71: Aqueous Solutions at Elevated Termpatures 3121-3136. Helgeson, H.C. 1968. "Evaluation of Irreversible Reactions in Geochemical Processes Involving Minerals and Aqueous Solutions I. " Geochim. Cosmochim. Acta, 32: 835-857. ‘Thermodynamic Relations. Helgeson, H.C. 1969. "Thermodynamics of Hydrothermal Systems at Elevated Tenperatures and Pressures." Amer. J. Sci. 267: 729-804. Helgeson, H. C. 1970. “A Chemical and Thermodynamic Model of Ore Deposition in Hydrothermal Systems." In Special Paper No. 3 Mineralogical Society of America 50th Anniv. Symposium, pp. 155-186, B. A. Morgan, ed., Mineralogical Society of America, Washington, D.C. Helgeson, H. C., Garrels, R. M., and Mackenzie, F. T. 1969. “evaluation of Irreversible Reactions in Geochemical Processes Involving Aqueous Solutions." Geochim. Cosmochim. Acta 33: 455-481. Helgeson, H. C., Brow, T. H., Nigrini, A., and Jones, T. A. 1970. "calculation of Mass Transfer in Geochemical Processes Involving 569-592. Aqueous Solutions." Geochim. Cosmochim. Acta 34: Welgeson, H. C., Delaney, J. M., Nesbitt, H. W., and Bird, D. K. 1978, “Summary and Critique of the Thernodynaaic Properties of 1-229. Rock-Forming Minerals." Amer. Jour. Sci. 278-1 Helgeson, H. C., And Aagaard, P. 1979. "A Retroactive Clock for Geochemical Processes". Abstract, Geol. Soc. Amer. Abstracts with Programs 11: 442. Jenne, E. A, ed. 1979. Chemical Modeling in Aqueous Systems. ACS Symp. Series 93, Anerican Chemical Society, Washington, D.C. Langmuir, D. 1978. “Uranium Solution-Mineral Equilibria at Low ‘Temperatures With Applications to Sedimentary Ore Deposits.” Geochim. Cosmochim. Acta 42: 547-569. Malinin, S. D., and Khitarov, N. I. 1969. "Reduction of Sulfate Sulfur by Hydrogen under Hydrothermal Conditions." Geochim. Int. 6 1022-1027. Nordstrom, D. K., et al. 19792. "A Comparison of Computerized Chemical In Models for Equilibrium Calculations in Aqueous Systems. Chemical Modeling in Aqueous Systems pp. 857-892, E. A. Jenne, ed., ACS Symp. Series 93, American Chemical Society, Washington, D.C. Nordstrom, D. K., Jenne, E. A. and Ball, J. W. 1979. "Redox Equilibria of Iron in Acid Mine Waters." In Chemical Modeling in Aqueous Systems pp. 51 - 59, Es As Jenne, ed., ACS Symp. Series 93, American Chemical Society, Washington, D.C. Ringwood, A. E. 1978. Safe Disposal of High Level Nuclear Wastes: A New Strategy. Australian National University Press, Canberra, Australia. Taylor, R. W., Jackson, D. D., Wolery, T. J. and Apps, J- A. 1978. ‘Section 5: Geochemistry." In Geothermal Resource and Reservoir Investigation of U. S. Bureau of Reclamation Leaseholds at Fast Mesa, Imperial Valley, Calfiornia, pp. 165-233. LBL-7094, Lawrence Berkeley Laboratory, Berkeley, California. Thorstenson, D. C. 1970. "Equilibrium Distribution of Small Organic Molecules in Natural Waters." Geochim. Cosmochim. Acta 34, ¢ 745-770. Wolery, T. J. 1978. Some Chemical Aspects of Hydrothermal Processes at Mid-Oceanic Ridges-A Theoretical Study. Ph.D. thesis, Northwestern University, Evanston, Ill- Wolery, T. J. 1979. Calculation of Chemical Equilibrium Between Aqueous Solutions and Minerals: The EQ3/EQ6 Software Package. UCRL-52658, Lawrence Livermore Laboratory, Livermore, California. Wolery, T. J. 1979b. "Sea Water-Oc Some Theoretical Considerations 1 Crust Hydrothermal chenistrys "abstract, EOS 60; 863. Wolery, T. J. 1980. Chemical Modeling of Geologic Disposal of Nuclear Waste: Progress Report and a Perspective. UCRL-52748, Lawrence Livermore National Laboratory, Livermore, California. jork performed under the auspices of the U.S. Department of Energy by the Lawrence Livermore National Laboratory under contract number W-7405-ENG~48." This document was prepared as an account of work sponsored by an agency of the United States Government. Neither the United States Government nor the University of California nor any of their employees, makes any warranty, express or implied, or assumes any legal liability or responsibility for the accuracy, completeness, or usefulness of any information, apparatus, product, or process disclosed, or represents that its use would not infringe privately owed rights. Reference herein to any specific commercial products, process, or service by trade name, trademark, manufacturer, or otherwise, does not necessarily constitute or imply its endorsement, recommendation, or favoring by the United States Government or the University of California. The views and opinions of authors expressed herein do not necessarily state or reflect those of the United States Government thereof, and shall not be used for advertising or product endorsenent purposes.

You might also like