Hydrogeology of the Lower Cretaceous Edwards and Trinity Group Formations Junction (Kimble County) Texas near by Stephen Robert Allen, B.S. Thesis Presented to the Faculty of the Graduate School of The University of Texas at Austin in Partial Fulfillment of the Requirements for the Degree of Master of Arts THE UNIVERSITY OF TEXAS AT AUSTIN May, 1997 Hydrogeology of the Lower Cretaceous Edwards and Trinity Group Formations Junction (Kimble County) Texas near Acknowledgments I wish to express my appreciation to the many individuals who provided guidance and support to during this degree program. your own me You have all in special way served as facilitators in the unfoldment of my potential. David Hill, Jun and Bruce Liao, Darling checked my answers to homework problems during the coursework phase. Jim Mayer, Jack Sharp, Clark Wilson, and Norman Van Broekhoven helped to clarify field interpretations of structure and stratigraphy. Philip Bennett allowed me to use lab to his ion chromatography perform analyses, and Todd Minehardt assisted by setting up the 1C equipment and explaining the use of standards. Alan Fryar and lan Jones answered questions I had on interpreting chemistry results. Leo Lynch volunteered his help by x-raying sediment samples to determine clay content. Barry Hibbs worked with me to get the numerical model set up and calibrated, and Robert Mace discussed how he resolved problems he encountered while models. Erica checked the using ground-water flow Boghici of borehole log cross-sections. was able to include color accuracy my I graphics in the thesis because I learned to use several software packages under the guidance of Reuben Reyes and Koko Kishi. Clark Wilson, Joel Davidow, Richard Weiland, and Tor Steineke edited the thesis. The final thesis review was performed mainly by Barry Hibbs who directed the research on a day to day basis. Larry Land, Peter Rose, Rene Barker, Jack Sharp, and Keith Young evaluated the proposed thesis goals and scope of work before activities were started. My thanks go to the individual landowners in the Junction, Texas area who allowed me to collect data on their properties. I’m especially indebted to 0. C Fisher, former U. S. Congressman who allowed me to stay at his vacation house near London, Texas while 1 worked in the field. III Abstract Hydrogeology of the Lower Cretaceous Edwards and Trinity Group Formations near Junction (Kimble County) Texas by Stephen Robert Allen, M.A. The University of Texas at Austin, 1997 Co-supervisors: Clark R. Wilson, Barry J. Hibbs This study describes ground-water flow in the Lower Cretaceous formations Texas. Rock near Junction (Kimble County), exposures were examined throughout the 150-mile study area to determine the nature and distribution of permeable features. Dominant features include nearly vertical fractures and horizontal bedding planes in carbonate rocks of the Edwards Group formations (Edwards), and coarse- grained fluvial channel deposits in the underlying Hensel Sand Formation (Hensei). Static water levels were measured in over one hundred wells and contoured to reveal the existence of two separate potentiometric surfaces, one overlying the other. but useful estimates of and Preliminary, transmissivity, hydraulic conductivity, ground-water velocity were derived using specific capacities from eighty-three wells completed in both the upper (Edwards) and lower (Hensel) aquifers. At the edge of the Edwards Plateau where the contact between the Edwards and the Hensel is exposed on the face of the erosional escarpment, ground water discharges from the Edwards aquifer through numerous low volume springs IV leaks the Edwards aquifer to the underlying Hensel aquifer across a thin low permeability bed at the base of the Edwards which consists of marly, unfractured, nodular and seeps. An even greater proportion of ground-water discharge from limestone. The sum of these two components of discharge is approximately equal to precipitation recharge to the Edwards. To gain additional insight into cross-formationai flow an analysis of major and minor ions, redox potential, and dissolved oxygen was conducted for twenty-one water wells which were located along three north-south (inferred flow direction) transects. The Edwards waters were found to be a Ca-Mg-HC03 facies; the Hensel waters, a mixed facies. This difference in hydrochemical facies was initially thought to be caused by ion evolution along flowpaths, but it more likely reflects the existence of a regional aquifer below a locally constrained aquifer. High values of dissolved oxygen and redox potential in the Edwards aquifer indicate that recharge is predominant; lower values of these parameters in the Hensel aquifer indicate that this water occurs in an intermediate or discharge zone. To test the conceptual model of steady-state ground-water flow, a numerical model was constructed MODFLOW finite-difference using the computer code. Over one hundred trial and simulations were executed to calculate error leakage through the confining bed, discharge from springs, and discharge to the Llano River. In addition, the distribution and magnitude of focused recharge to the Edwards aquifer was calculated, estimates were made for unknown hydrogeologic parameters, and the Edwards aquifer was demonstrated to be fully perched above the Hensel aquifer. The increased understanding of the ground-water flow regime resulting from this study will support range management activities and improve the success rate of water well drilling. V Table of Contents Acknowledgments iii Abstract iv Table of Contents vi Introduction and Geology 1 Methods and Data 36 Results of Field Observations and Water Chemistry Analyses 43 Conceptual Model of Ground-Water Flow 73 Mathematical Model of Ground-Water Flow 81 Conclusions 110 Appendices 114 Appendix A Water chemistry sampling and analysis procedures 115 119 Appendix B Rainfall record Appendix C Descriptions of soil types 121 Table of water levels in wells 123 Appendix D Appendix E Specific capacity and transmissivity calculations 128 Appendix F Aquifer recharge calculation 132 Appendix G Table of water chemistry values 136 Bibliography 145 Vita 155 VI Introduction and Geology Research objectives and scope of work a This study describes the ground-water flow regime in 390 square km (150­ square mile) area at the edge of the Edwards Plateau in Kimble County, Texas. Along this escarpment the Lower Cretaceous Fort Terrett and Segovia Formations of the Edwards Group (Edwards) and the underlying Hensel Sand Formation (Hensel) of the Trinity Group are exposed. The study resulted in the development of a conceptual model which describes recharge, discharge, head-dependent flux through a thin low-permeability bed between two aquifers, and flow within the saturated zone. controlled The model explains how hydrogeologic inputs, throughputs, and outputs are by stratigraphy, structure, and topography. Fieldwork consisted, first, in measuring the water levels in 112 wells to produce potentiometric surface maps. Next, the water from twenty-one of the wells These two was analyzed for major ions, minor ions, unstable indicator parameters. sources of data guided the mapping of recharge and discharge areas, and supported the hypothesis of two separate aquifers. The finite-difference program MODFLOW (McDonald and Harbaugh, 1984) was used to create a quasi three-dimensional, steady-state model for the purposes of visualizing the ground-water flow regime, testing the validity of assumptions used to construct the conceptual model, and verifying and adjusting estimated hydrogeologic parameter values. 1 Location and physiography The study area lies between north latitudes 30° 30’ and 30° 42’ 30”, and west latitudes 99° 35’ and 99° 50’, at the eastern edge of the Edwards Plateau in Kimble County, Texas (Figure 1). It is located 210 km (130 miles) west of Austin and 177 km (110 miles) northwest of San Antonio. Surface geology and major physiographic features are shown in Figure 2. The edge of the plateau is represented by erosional escarpment having average of approximately an an relief 60 meters (200 feet) (Figures 3 and 4). This escarpment extends across the center of the Its existence has study area and strongly influences the ground-water flow regime. the effect of separating the study area into two distinct physiographic regions: a highly dissected, horizontally bedded carbonate upland composed of the Lower Cretaceous Edwards (Figures 5 and 6); and, the Llano River floodplain, which is veneered with alluvium atop the siltstone, sandstone, and conglomerate of the Lower Cretaceous Hensel (Figure 7). Relief from 713 meters (2,340 feet) at the ranges topographic divide to 475 meters (1,560 feet) at the lower (eastern) end of the Llano River. Significance of the study This study investigates the physical and chemical hydrogeology in the Junction, Texas area. It provides a framework from which to understand cross­ formational flow and the origin and distribution of solutes. The increased knowledge of aquifer quality, storage and capacity can also be used by landowners to exploit this limited resource more effectively. Physical hydrogeology Two separate potentiometric surfaces were drawn from water levels measured in wells that were located throughout the study area. These potentiometric will establish depths to the surface maps, when overlayed by the topographic map, 2 Figure 1. Map of regional geology in central Texas. The study area is located in part, at the dissected edge of the Edwards Plateau; and in part, on the Llano River floodplain (modified from American Association of Petroleum Geologists, 1973). 3 Figure 2. Base map of the study area showing surface geology, watersheds, the Edwards Plateau escarpment, and the Llano River floodplain. The exact distribution of alluvium in the study area has not been established (modified from Barnes, 1981). 4 Figures. Topography in the study area. Elevations range from 475 meters (1,560 feet) at the eastern end of the Llano River to 713 meters (2,340 feet) at the highest topographic divide (modified from Barnes, 1981). 5 area study mile 150-square the of Physiography 4. Figure Figure 5. Boundary between the flat top of the Edwards Plateau and one of the many deeply incised tributaries that originate at the edge of the plateau. 7 ephemeral surface. Plateau. for upland flat outlets Edwards its as via the serve of edgerocks Formations at the carbonate Group head its in the Edwards to close planes the valley bedding infiltrated tributary horizontal has that and steep-walled rainwater A of fractures 6. Figure Vertical discharge the background is along here in floodplain the The (seen looms Texas. alluvium Plateau Junction, with Edwardsveneered the near is of River which Llano escarpment Formation underfit erosional the Sand of The HenselFloodplain river). the bythe 7. of Figure underlain banks saturated zone at potential drilling locations. In the past, drillers have not understood the nature of ground-water occurrence in this region. The relationship between relief and water-table depth in the highly dissected region at the edge of the Edwards has not been recognized. This has resulted in the drilling of non-productive wells in places where the saturated zone is thin or non-existent. Numerical flow simulation assisted in evaluating the distribution of recharge. Spatially variable recharge was assigned to the model grid, and the simulated water-table contours were compared to hand-drawn contours. The recharge pattern leading to the best match was used to calibrate the model and to derive estimates of rate and spatial variability of recharge. Observations of springs and seeps, and aerial photos of tree line patterns were used to identify areas of ground-water discharge. A comparison of present and historical tree coverage (as seen in photographs provided by landowners) shows that an increase in the number of mesquite and juniper trees since earlier in the century correlates with a reduction in ground-water discharge from Edwards aquifer springs. The spread of opportunistic vegetation over time (Figure 8), and the related loss of grass cover, are likely to be responsible for a reduction in aquifer recharge, and a concomitant reduction in spring discharge. This observation suggests it be of may benefit to remove these species of trees to increase the throughput of water. A study by Dugas and Hicks (1994) in the Seco Creek watershed in Texas concluded that spring discharge increased after junipers were removed from a selected area. Chemical hydrogeology The analysis of major ions and indicator parameters (DO, pH, Eh, a water temperature, and alkalinity as bicarbonate) in twenty-one wells establishes baseline. lon concentration data were used to determine the quality spatial distribution of hydrochemical facies, and to map the locations of aquifer recharge and discharge zones. Variations of hydrochemical facies correlated mostly to transitions Juniper and Mesquite as such species Opportunistic 8. Figure century. the in earlier since coverage spatial in increased have in lithofacies and, to a much lesser extent, to hydrochemical evolution along flowpaths. Environmental setting Land use and economics The of economy the study area is based primarily on ranching, outdoor recreation, and interstate travel. Limited withdrawals are made from the Llano River to and irrigate hay fields and pecan orchards near the river. Deer, turkey, numerous exotic species graze on private property where they are harvested by hunters who lease the land. Ground water contributes directly to the as economy water supply and, indirectly, as a scenic feature, with springs and spring-fed streams contributing to the recreational value of the area. Climate and weather The study area is located in a subtropical subhumid climatic region characterized by hot humid summers and dry winters (Larkin and Bomar, 1983). Onshore flow of tropical maritime air from the Gulf of Mexico is the dominant control of climate. Intermittent seasonal intrusions of continental exert regional air, secondary influences on the climate. Droughts and floods occur in Texas on a regular basis. Floods are caused by thunderstorms in the spring and tropical storms in late summer. Spring thunderstorms occur along the line of contact between cold fronts and overtopping warm, moist air from the Gulf of Mexico. In late summer, tropical cyclones originate in weather systems that have their beginnings in the Caribbean Sea or the Gulf of Mexico. Droughts are caused mainly by the extensive subtropical high pressure cell (the Bermuda High) that drifts latitudinally with the passing of the seasons. When the Bermuda High is entrenched over the southern United States the possibility of drought becomes more likely (U. S. Geological Survey, 1989). In the Junction, Texas area the average annual precipitation for the period 1939-1993 (Dunk, 1994; Appendix A) 61.34 cm was (24.15 inches). Two graphs of precipitation are shown in Figure 9. One shows the average monthly precipitation for this period; and the other, the total rainfall for each year in the period of record. Monthly rainfall maxima occur in late spring and early fall; minima occur in midwinter and early summer (Dunk, 1994). The monthly gross lake evaporation rate during the period 1951-1980 ranged from an average of 6.60 cm (2.60 inches) during January to an average of 25.40 cm (10.0 inches) during August for an annual average of 177.80 cm (70.0 inches) (Larkin and 80mar,1983). Vegetation A variety of vegetation is present on the Edwards Plateau. Prairie grasses, live oak and Spanish oak, and "cedar" (scrub juniper) grow on the limestone upland and marly dissected zones. Lining the banks of the creeks and rivers are cypress trees. Terraces live and support growths of post oak, juniper, elm, hackberry, cottonwood, sycamore, and willow. Natural grasses include little bluestem, Indian grass, sideoats grama, and Texas winter grass. Introduced grasses include coastal Bermuda, plains lovegrass, Klein grass, and King ranch bluestem (Cuyler, 1931). Trees, such as the juniper, are wasteful water compared to grasses users because they release moisture to the atmosphere through their stomatas twenty-four hours a day. Some landowners report that where junipers have been removed, It creeks and springs on their property have increased in discharge. is hypothesized that excessive growth of juniper and other opportunistic trees have caused a reduction in aquifer recharge, and a concomitant reduction in spring discharge. Figure 9. Histograms of rainfall in Junction, Texas from 1939 to 1991. The upper histogram shows average monthly rainfall; the lower, total annual rainfall (Dunk, 1994). 14 Soils Soils in the study area are divided into four types by the Soil Conservation Service (SCS) (Blum, 1982). This division is based on sets of physical properties that control land use. in soil Mapped Figure 10, the types in the study area areTarrant, Tarrant-Real-Brackett, Nuvalde-Dev-Frio, and Menard-Hext-Latom. Detailed descriptions are included in Appendix C. During the initial phase of this study, the use of soil thicknesses and infiltration rates reported by the SCS was considered as an approach for estimating the spatial distribution of recharge because these factors are correlated with rate of recharge. This approach was discarded, however, because other factors such as relief, structure, and stratigraphy were found to be more important. In addition, Chock Woodruff (pers. comm., 1995) performed field experiments which led to the conclusion that soil thicknesses recorded by the SCS for the Edwards Plateau-Hill Country area were underestimated because they had been measured with handtools rather than backhoes. These underestimates of soil thicknesses could lead to underestimates of aquifer recharge because thinner soils would likely have a lower infiltration capacity that would result in higher runoff of rainfall. Geology The surface geology of the study area (Figure 2, Barnes, 1981), is organized by relative age in the stratigraphic column (Figure 11; Maclay and Land, 1984). The the Edwards Plateau escarpment (Figure 12) exposes stratigraphy represented in Figures 2 and 11. upon a topographically rugged, pre-Cretaceous These strata rest unconformity named as the Wichita Paleoplain (Hill, 1898). Strata were deposited in conjunction with the eustatic rise and northwestward transgression of the Lower Cretaceous Comanchean Sea. Referring to Figure 12, the lower strata is the Hensei Sand Formation (Hensei) of the Trinity Group, and the upper is the Fort Terrett Figure 10. Distribution of soil types in the study area (Blum, 1982). Figure 11. Generalized stratigraphic column in the Junction, Texas area (modified from Maclay and Land, 1984). 12. Figure Stratigraphy on the face of the erosionai escarpment of the Edwards Plateau. Exposed from bottom to top is the red-orange, friable siltstone at the top of the Hensel Sand Formation, the basal marly limestone member of the Fort Terrett Formation, and the fractured limestones and dolomites of the Fort Terrett Formation of the Edwards Group (Edwards). The Fort Terrett Formation and the overlying Segovia Formation of the Edwards Group are together referred to in this report as the Edwards except in cases where it is important to distinguish between the two. The aquifer occuring in both formations will be referred to as the Edwards aquifer. Hensel Sand Formation The Hensel, which is Upper Trinity in age, was deposited contemporaneously with the downdip (southeast and east of the study area) Glen Rose Formation. “This is the of three clastic-carbonate depositional couplet youngest couplets, that separated by disconformities, reflect a pattern of cyclic sedimentation superimposed on an overall transgressive regimen” (Stricklin and others, 1971). Each couplet generally onlaps rocks of the previous cycle and documents a major advance of the early Cretaceous sea terminated by an overall drop in sea level. “Episodic rejuvenation in the source area resulted in an increased supply of elastics detrital followed and a consequent depositional phase, by relatively quiescent sedimentation of carbonate deposits, the latter phase in part contemporaneous with and transitional into the clastic phase” (Boone, 1968). During Jurassic time (60 million year period) most of west-central Texas was emergent and subject to erosion prior to inundation by the Comanchean Sea. Extensive northwest-southeast had been carved out of trending river valleys that the Paleozoic surface, became filled with terrigenous Trinitian sediments to age cause a substantial leveling of the ground surface. Eventually, these sediments formed a wedge which thickens downdip (toward the southeast). The Hensel sediments which fill these valleys were derived from erosion of the Paleozoic surface, and from erosion of Precambrian rocks that were exposed as a paleohigh in In the the thickness of the Hensel is the Llano Uplift region. study area average approximately 60 meters (200 feet). cross-sectional view An interpretation of the Hensel paleoenvironment and a of the of Lower is sequence depositional events occuring during the Cretaceous Previous studies depicted in Figure 13 (Payne, 1982). (Rapp, 1988; Payne, 1982; Barnes, 1981; Hall, and Turk, 1975; Amsbury, and others, 1974; Inden, 1974; and Stricklin, others, 1971; Boone, 1968; Campbell, 1962; Hughes, 1948; Hill, and Vaughan, 1898) have established the Hensel to be a its cemented conglomerate at base which fines upward to become a friable calcareous siltstone (Figure 12) at its top. The basal conglomerate is attributed to initial high stream gradients, and the fine-grained material at the top is attributed to the low stream energy that developed as the relief became leveled (Stricklin and others, 1971). Sedimentary features observed in the field such as thin beds of carbonate, cross-bedding, ripple marks, crevasse splays, oyster shell orientations, psuedoanticlines, and caliche have lead to arid to semi-arid The Hensel is the interpretation of an fluvial paleoenvironment. described by Hall and Turk (1975) as a dip-oriented multilateral sandstone body characterized by two important facies: (1) a meanderbelt sandstone facies floodbasin facies (lenticular coarse-grained framework deposits), and (2) a (finer­grained matrix deposits). architecture within the Hensel results in The complex facies an equally of a that complex pattern ground-water flow. An example of diagenetic feature contributes to this complexity is presented in Figure 14. This is a photograph of an with extensive nodular caliche that was deposited between individual grains exposure in the sandstone by circulating ground water. This caliche reduces primary porosity and probably increases the tortuosity of ground-water flow. Edwards Group Formations During the Lower Cretaceous Period the study area was situated approximately 25° north of the equator within the belt of prevailing easterlies. at around 30° north resulted in dry adiabatic warming and Descending air masses Figure 13. Fluvial paleoenvironment (upper diagram) and model (lower diagram) of Trinity Group deposition (from Payne, 1982). out oxidized precipitated The has Formation. that Sand carbonate)Hensel the (calcium of part caliche middle with the in Outcrop impregnated is 14. groundwater. Figure silty-sand of low relative humidity (Rose, 1972). Rainfall, possibly associated with tropical storms was distinctly seasonal. This climate was conducive to the deposition of the carbonates and that formed the Edwards These evaporites Group. sedimentary rocks were deposited in shallow water on the broad Comanche Shelf whose dominant feature the study area) was the Central Texas Platform, a wide (in elongate positive surface that was leveled by deposition of terrigenous Hensel sediments that filled pronounced valleys on the Paleozoic (Wichita Paleoplain) surface (Figure 15). The Edwards is characterized rudist bioherms, carbonate and by grainstones and down mudstones, evaporites laid during three transgressive-regressive sequences under predominantly shallow marine conditions of relatively low wave and current energy (Rose, 1972). These relatively thin, nearly flat-lying strata (Figure 16) dip gently southeastward atop massive Paleozoic and Triassic units that dip westward. By early Fredricksburg time, an offshore bioherm of rudists, corals and carbonate deposits, named the Stuart City reef trend, had grown to an almost continuous ridge along the seaward edge of the continental shelf in the ancestral Gulf of Mexico basin (Figure 17) (Bebout and Loucks, 1974; Fisher and Rodda, 1969; Rose, from 1968). The rapid growth of this reef trend probably resulted a rapid rise in sea level that may have been triggered by an increase in the rate of seafloor spreading (Bay, 1977). This reef trend sheltered intertidal and restricted marine depositional environments on its leeward side (Comanche Shelf) from deep, open marine conditions in the ancestral Gulf of Mexico basin. During periods of especially low sea level and extreme aridity, the crest of the central Texas Platform became a broad, sahbka-type mudflat where evaporites, dolostone, and thin-bedded dolomitic limestone Sedimentation was controlled by several were deposited (Fisher and Rodda, 1969). factors; climate, influx of terrigenous clastic sediment, distribution of tectonic subsidence and uplift, and energy level of wave and current action (Barker, and 15. Tectonic framework of central Texas. Cretaceous Figure age carbonate rocks of the Edwards Group were deposited in a broad shallow sea named the Comanche Shelf (from Rose, 1972). undisturbed and thickness,relatively lithology, its environment. Note texture,Group. the depositional Edwards in the transitions the in Vertical of adjustments Formation Segovia horizontality. periodic the of reflect of degreebeds Outcrop high individual 16. and of Figure nature color Figure 17. Lower Cretaceous paleoenvironment in central Texas (from Fisher and Rodda, 1969). others, 1994). The resulting tithofacies determined the stratigraphy, and together with the effects of subsequent tectonics and diagenesis, the hydraulic characteristics of rocks that today compose the Edwards aquifer area the Edwards two the In the study Group is composed of formations, Fort Terrett and the overlying Segovia. Rose (1972) elevated the Edwards from formation to group status, the Fort Terrett and Segovia from member to formation status, and several informal members to member status. These rank changes were made because the Fort Terrett and Segovia each contain significant rock units that should be called members because of their thickness and vertical variability. as Accordingly, the units comprising them rank formations constituting yet a larger unit, the Edwards Group. The base of the Edwards below the Fort Terrett is a probably disconformity, the rock above being distinctly more marine. For purposes defining of is drawn at ground-water flowpath boundaries, the base of the Edwards the change upward from recessive argillaceous rock to resistant massive limestone. The Fort Terrett, which is approximately 58 meters (190 feet) thick in the Junction area, was deposited in environments ranging from open marine to intertidal, all with low wave and current Members composing the Fort Terrett are energy. from bottom to top, the Basal Nodular Member, Burrowed Member, Dolomitic Member, and Kirschberg Evaporite Member. These are described by Rose (1972) as follows: Basal Nodular Member-A nodular marly zone up to eight meters Exogyra texana, lunatid and turritellid (twenty-five feet) thick, rich in snails, and protocardid clams. Burrowed Member-Massive, resistant layers of porous, burrowed limestone approximately twenty-five meters (eighty feet) thick, thin beds of miliolid and mollusc-fragment biosparite and dolomite, deposited in a restricted to open shallow marine environment. This member is the chief water-bearing zone of the Edwards, its porosity caused in part by preferential leaching and removal of burrow-fillings to produce widespread honeycomb porosity. Dolomitic Member-Massive to thin-bedded, fine to medium crystalline, homogeneous dolomite, with abundant chert and beds of miliolids and rudistid biosparite, deposited in an intertidal to restricted shallow marine environment. Klrschberg Evaporite-Widespread, twelve to twenty meter (forty to eighty foot) thick disturbed and altered zone that marks the former horizon. presence of a gypsum Massive to thin-bedded, cherty, crystalline limestone and travertine containing intervals of limestone, dolomite and chert fragment collapse breccia. Its depositional environment was restricted shallow marine, and in part sahbka-type supratidal and intertidal (Figure 17). The Segovia, which is approximately 35 meters (115 feet) thick in the Junction area, was deposited in environments ranging from marine to intertidal. open It is composed of the Burt Ranch Member, Alien Ranch Breccia Member, Orr Ranch Bed Member, and Black Bed Member. In the Junction area the Orr Ranch Bed and does notthe Black Bed are eroded atop the Segovia, and the Allen Ranch Breccia extend this far north and west. Unlike the Fort Terrett, which is divided into four intergradational members, the Segovia is subdivided into members on the basis of thin, distinctive widespread key beds. The section in the Junction area is described by Rose (1972) as follows: Burt Ranch Member-Persistent and widespread zone of marly limestone, above the Kirschberg Evaporite Member having a thickness of approximately twenty meters (seventy feet). It includes marl, marly micrite, miliolid biosparite, and rudist biosparite with a few scattered beds of soft massive dolomite. Clayey or marly fossiliferous zones are particularly common near the base and toward the top and contain Exogyra texana, and turritellid snails. This member reflects distinctly more marine depositional conditions than the underlying Fort Terrett. This claim is supported by its rich and diverse mollusk and ammonite fauna which indicates deposition on a shallow shelf. open Unnamed Member-Atop the Burt Ranch Member in the Junction area is an unnamed thin-bedded limestone biomicrite and micrite with isolated beds of dolomite about 180 meters (55 feet) thick. Deposition of carbonates continued during Washita time, but over most of the Plateau, as well as in the study area these sediments are now eroded. Toward the end of the Washita, in response to an upwarping of the Comanchean Shelf, there was a widespread regression of the shallow sea. The Washita Group strata was stripped away and the Edwards became exposed. Since becoming exposed, the Edwards has alteration such undergone diagenetic by processes as fracturing, recrystallization, cementation, dissolution, and collapse of These resistent beds. processes have produced variations in porosity and permeability both laterally and vertically, which have influenced aquifer heterogeneity. Alluvium and Colluvium area. Several types of Quaternary surficial deposits are found in the study These are classified as either one of several types of alluvium or as colluvium (Barnes, 1981). Deposition of these sediments occured from the Pleistocene to the Recent, but their exact ages and modes of formation are not known because they have not been studied (Michael Blum; pers. comm., 1995). Close to the Llano River, near the level of the present floodplain, is a ten (33 foot) layer of (Figure 7) some places supports meter thick alluvium that in extent. Three of alluvial shallow, potable aquifers of limited types deposits were mapped by Barnes (1981) using aerial photos: (1) those that formed in Recent times terraces along the sides of tributaries which are perpendicular to the Llano River, (2) deposited in Recent times which parallel the present course of the Llano River and, (3) remnants of high terraces flanking the edges of the underfit Llano River valley that are older than the low terraces. The distribution of alluvial deposits is represented in a on generallized sense the map of surface geology (Figure 2). A photograph of a coarse-grained point bar deposit in a tributary that drains the dissected area next to the Edwards Plateau is shown in Figure 18. These rounded gravels which are composed of limestone, dolomite, and chert, were deposited during intense, short duration floods. The source of most of the gravel is probably colluvium that has accumulated at the bottom of terrace steep slopes of the Edwards higher in the tributary reaches. Exposures of gravels throughout the area reveal that they are constructed of a mixture of well- rounded chert and limestone A gravel loosely bound by fine-grained travertine. present day example of fine-grained deposition within the interstices of gravel deposits is shown in Figure 19. This is a photograph of the bed of the Llano River showing how calcium carbonate (travertine) precipitates out of solution when carbon dioxide is lost from the river water. The calcium carbonate settles into the intergranular spaces of the limestone and chert gravel. The result is a tightly-bound alluvial deposit having a bimodal distribution of grains and perhaps a diminished infiltration capacity. Throughout the area, near the transition between the Edwards and the occur. Hensel, widespread fanplains These surface deposits were referred to as point across These ran that tributary. waters flood ephemeral by deposited meandering bar point a within Coarse-grained floodplain River 18. Llano Figure the downward. cuts stream the as terraces form to time over up build bars Figure 19. Bed of the Llano River showing a fine-grained travertine (calcium carbonate) ooze building up between the pebbles and cobbles. Travertine is precipitated from the river water when carbon dioxide is lost to the atmosphere. 32 “washes” by Hill (1898). Apparently, they are deposited by overland runoff that contains an unsorted mixture of colluvium. Previous investigations Previous investigations are grouped into two categories: (1) local hydrogeologic and surface hydrology studies, and (2) hydrogeologic studies outside of the local area yet within the same physiographic province. Local hydrogeologic and surface hydrology studies The ground-water study that relates most substantially to the present one is by Alexander and Patman (1969). They “determined the occurrence, availability, dependability,quality, and quantity of the ground-water resources of the County” and provided a qualitative overview of hydrogeology and surface-water hydrology. A stream gauging study was completed by Holland and Mendieta (1965). From January 17 to January 24, 1962, the flow rates of the Llano River and its principal tributaries were measured at fifty-three points between Junction and Llano, Texas with a portable stream gauge. Holland and Mendieta believed that the flow of the Llano River and its tributaries was being sustained contact No by springs. runoff-producing rains had occured for sixty-six days prior to the investigation. Based on this observation they determined that measured discharge values would low-flow conditions. It quantify was hoped that their results would shed light on interaction to the surface-water/ground-water support present study but, unfortunately, because the rate of discharge measured from one stream to gauge the next did not differ by more than five percent, the data could not be used. Five percent is the margin of error for stream gauging with the portable Price current meter. Hydrogeologic studies outside of the local area vet within the same physiographic province A number of hydrogeological studies within the Edwards or Hensel in the Edwards Plateau/Hill Country region were completed outside the study area. These provided insight into the hydrologic regime of the study area, and contained data that were compared with those of this study. These data included transmissivities, hydraulic conductivities, thicknesses, formation thicknesses, spring outlet saturated elevations, well yields, major ion values, and fracture orientations. Previous studies include those of: (1) Barker and others (1994) who provided an up-to-date overview of the hydrogeology of the Edwards-Trinity aquifer by condensing an array of previous studies. (2) Kuniansky and Holligan (1993) who presented the results of a digital model of the ground-water flow system in the Edwards-Trinity aquifer. They calculated water budgets for pre-development and post-development conditions, and found that ground water development reduced spring flow and leakage to streams. (3) The Hill Country Underground Water Conservation District (1994) who produced the "Gillespie County Regional Water Management Plan." (4) Bush and others (1993) who presented a three-sheet map series on the historical surface the These potentiometric of Edwards-Trinity Aquifer System. maps portray regional water-level patterns over broad areas using the earliest available data at 1,789 well locations. (5) Ardis and Barker (1993) who created a two-sheet map series illustrating the saturated thickness of the Edwards-Trinity Aquifer System and selected contiguous hydraulically connected units in west-central Texas. (6) Barker and Ardis (1992) who reported on the configuration of the base of the Edwards-Trinity Aquifer System and hydrogeology of the underlying pre-Cretaceous rocks, west-central Texas. They reported that the Cretaceous aquifer system is underlain by an extensive complex of rocks ranging from Late Cambrian through Late Triassic in that are typically ten to one thousand times less age permeable than those comprising the aquifer system. (7) Bluntzer (1992) who evaluated the ground-water resources of the Paleozoic and Cretaceous aquifers in the Hill Country of Central Texas. He claimed that the projected population increase between 1985 and 2010 would double the demand for water from 22,872 acre-feet/year to 47,380 acre-feet/year. (8) Abbott and Woodruff (1986) who edited a compendium of papers that addressed the geology, hydrology, ecology, and social development in Central Texas near the the is the Balcones Escarpment. Particularly relevant to present study paper by the Edwards Rose (1986) concerning the potential impact of pipeline oil spills upon Plateau Aquifer. He concluded that the aquifer is exceptionally vulnerable to pollution from pipeline oil spills because its fracture permeability would allow oil to sink into the bedrock before cleanup crews could respond. Ashworth who assessed the of the Lower (9) (1983) ground-water availability Cretaceous formations in eleven counties in the Hill Country. Ashworth described the depositional environments, stratigraphy and structure of the Lower Cretaceous formations and how these factors control the hydrogeology. (10) Walker (1979) who reported on the occurance, availability, and chemical quality of ground water in the Edwards Plateau region of Texas. He defined five aquifers of different ages located in the twenty-eight county study area. Methods and Data Data used in this study came from three sources: existing records, field measurements, and laboratory analyses. Data collection and processing, and the methods used to generate new data are described. Existing records included Llano Existing records utilized drillers’ reports, specific capacities, River discharge rates, and aerial photos. Water well inventory One hundred and twelve water wells were inventoried using data from two State of Texas record archives (Appendices D). Well location maps and drillers’ reports for “plotted” wells were obtained from the Texas Natural Resources Conservation Commission (TNRCC). Well location maps and hydrochemical analysis values for “located” wells were obtained from the Texas Water Development Board (TWDB). Plotted wells well locations Maps of plotted wells show approximate on U. S. Geological Survey (USGS) 7.5-minute topographic maps, as reported to the TNRCC by individual drillers. These wells assigned approximate locations on Texas Department of Water well contain Transportation county highway maps. reports may information on specific capacity, static water level, lithology, well diameter, and well and screen depths. Located wells Located wells are field located by TWDB technicians on USGS 7.5 minute In the of these wells topographic maps. study area, many are easily spotted because use windmills for include information on static they pumps. Well reports water levels, and field-measured values of specific conductance, and total iron (Alexander and Patman,l969). Reported iron analyses are ferric values. Ferric iron is the undissolved form of iron contributed by rusty well hardware. Specific capacity tests Some well reports submitted to the TNRCC included the results of well acceptance tests. These data were used to calculate specific capacity, which in turn was used to estimate transmissivity and hydraulic conductivity (Appendix E). By pumping for a period of time and noting little or no drawdown of the water table, the driller “accepts” the well for its intended use. The well acceptance test is conducted for a much shorter duration (0.5 to 2.0 hours) than the aquifer pump tests For this performed by hydrogeologists to calculate hydrogeologic parameter values. study area, specific capacity calculations were the only source of data to make estimates of transmissivity and hydraulic conductivity. Fifty-seven specific capacity values from the Hensei aquifer and twenty-six from the Edwards aquifer were used to estimate transmissivity. Hydraulic was then calculated the conductivity by dividing transmissivity by average saturated thickness of each aquifer. The formula used was the modified non-leaky artesian formula (Walton, 1962): Q/s = T/(26410g(Tt/2693r2S)-65.5) (eq. 1) where: 2 Q/s = specific capacity (L /1) T = transmissivity (L2/t) S = storage coefficient [dimensionless] well radius (L) r= t = time since pumping started (t) This equation relates the specific capacity of the well to the transmissivity of the aquifer. It assumes that the well is discharging at a constant rate in a homogeneous, isotropic, nonleaky, artesian aquifer, infinite in areal extent. Estimates of S (.01 to .0001) bracketed the range of probable actual values because the degree of confinement within the portion of the Hensei aquifer occuring below the Edwards was not known. There are three reasons why specific capacity calculations must be considered approximations. First, pumping periods in some wells may be too short; thus, the water table is not given an adequate amount of time to become depressed as required by the equation. Second, well reports do not indicate how soon after the cessation of pumping that the static water level was measured. A delay between the cessation of pumping and measurement of the water level could cause an erroneously small drawdown which would result in the calculation of an overly large value of specific capacity. Third, the calculated average values of specific capacity probably underestimate T because test results in cases where zero drawdown is observed (i.e. higher transmissivities) cannot be processed by the Walton equation to calculate transmissivity. Llano River discharge records/precipitation records Discharge rates recorded at USGS stream gauge # 08150000 (Junction) were used to estimate baseflow in the Llano River. Data from the months of December and January for the years 1939 to 1993 were used (U. S. Geological Survey, 1994, and Appendix F) because they were low-flow months. Surface runoff and evaporation during low-flow months are at a minimum; thus, the water in the river is derived almost from baseflow. These data were with entirely coupled for the same months and for the precipitation values (Dunk, 1994, and Appendix B) same period of record to produce an estimate of annual ground-water recharge in the The estimated drainage basin upstream of the river gauge. recharge in cm/year was extrapolated to the study area and used as initial input to the numerical flow model. Low altitude aerial photographs Black and white aerial photographs of the study area were obtained from the Texas Natural Resources Information Service Distinct (TNRIS). patterns of tree growth observed on the photos correlate strongly with stratigraphic features that control spring discharge from the Edwards aquifer. New data acquired in the field and laboratory New data acquired included ground-water level measurements, dissolved ion concentrations, unstable constituent values, and results of an x-ray analysis of sediment. Ground-water level measurements Static water levels were measured in 112 wells (Appendix D). Interpolated data points were then used to construct potentiometric surface maps. Measurements were made in wells equipped with windmills and submersible pumps, and in wells with no installed Before measurements were made, the pump. pumps were turned off and the water table was allowed to recover. Depths to water were measured to the nearest .3 meter (one foot) with a chalked steel One foot tape. was considered sufficiently accurate because topographic relief in the study area is about 230 meters (750 feet) and ground elevation control is accurate only to about three meters (ten feet) (one-half of a contour interval) on 7.5-minute USGS topographic maps. Measurements were taken during an eight month period as access to each property could be secured. Repeat measurements of a half-dozen control wells were made the and water-level values were to within during study, replicated approximately one meter. To calculate the elevation of the potentiometric surface at each water well, the depth-to-water was subtracted from the ground surface elevation. Ground surface elevations were established from USGS 7.5-minute topographic maps, but where it was not possible to plot a well accurately on the an transit was used to from three topographic map engineering triangulate topographic features to determine location. Sampling and analysis of ground-water chemistry Ground-water samples were collected from twenty-one wells. Seven samples were taken from the Edwards aquifer, thirteen from the Hensel aquifer, and one from the thin alluvial aquifer next to the Llano River. Collection and analysis followed the set of procedures listed in Appendix A and diagrammed in Figure 20. To summarize, wells pumped until temperature, were dissolved oxygen, and pH achieved temporal stability. These indicator values were measured within a flowcell that isolated the aquifer water and the measuring probes from the atmosphere. Alkalinity was measured in the field by titrating with 1.5 N HCL to the 4.5 pH endpoint. Samples were collected in pre-cleaned polyvinyl chloride bottles and preserved for ion analysis in the laboratory. Anion concentrations were measured with the ion chromatograph (IC), and cation concentrations were measured using the inductively-coupled plasma spectrophotometer (ICR). Anion concentrations were determined using both the conductivity and the absorbance detectors and these calculations were then adjusted by the ratio of the !C measured value of a certified check standard to the manufacturer’s measured value. ionic charge balance to within 5.2 percent was achieved for each sample that HC0 N0 and N0 was analyzed for Ca, Mg, Sr, K, Li, Na, and Cl, F, S0 4 , 3, Br, 3 , 2 . An ionic charge balance error of less than five is believed to rule out percent Several constituents not measured in the lab were measured in the analytical error. field with a Milton Roy Mini-20 spectrophotometer. These included sulfide, dissolved Figure 20. Flowchart of procedures for the measurement of unstable constituents in the field, and collection of ground­water samples for major ion analysis in the laboratory. oxygen, total Fe, ferrous Fe, phosphate, and ammonia. Results of all water analyses are listed in Appendix G. X-ray diffraction analysis to classes Samples of fine-grained Hensel strata were x-rayed identify of clay minerals. Clay mineralogy is an important factor in controlling the major ion of water in sedimentary deposits due to the high surface area available chemistry for chemical exchange reactions. X-ray diffraction results are presented in the next chapter. Results of Field Observations and Water Chemistry Analyses Field observations Observations and measurements made in the field provide a basis for developing a conceptual hydrogeologic model of the study area. Field observations were made of topography, soils, streams, storm runoff, springs, vegetation, and rock outcrops. Measurements included static water levels, saturated thicknesses, major ion concentrations, and unstable constituent values. These measurements, and interpretations of existing data, were used to create maps of potentiometric surfaces, flowlines, recharge distribution, hydraulic conductivity distribution, redox These potential, hydrochemical facies distribution, and locations of contact springs. results were used up the finite-difference model of ground-water flow. to set Summaries of results follow. Topographic divide on top of the Edwards Plateau of the Edwards Plateau The relatively flat-lying upland surface (Segovia Formation) is a geomorphically old terrain in comparison to the highly dissected zone near the edge of the Plateau. Pictured in Figure 21 is the topographic (and hydrologic) divide at the higher elevation boundary of the study area. Years of overgrazing on open rangelands stripped grasses, soil erosion, and soil compaction. This created the type of landscape locally referred to as a “goatscape.” Surface water runoff is probably higher now than it was in the past, even though focused is to be on the flat surfaces of the recharge hypothesized prevalent plateau. Widespread shallow depression storage is observed to occur on this surface after storms, so it is likely that some of the trapped water percolates downward to at are estimated in the recharge the Edwards aquifer. Recharge rates the plateau mathematical flow model chapter. is Grazing This Group). erosion. (Edwards soil extensive to Formation aquifer. Segovia contributed the unconfined has of century upper surface for area Flat-lying mid-nineteenth the 21. the recharge Figure since the Fractures and bedding planes in the Edwards Group Formations Vertical fractures are observed throughout the Edwards Plateau (Figure 22). The dominant trend directions are approximately N4O°E and N4O°W. These match the dominant fracture trends inferred from lineaments mapped on aerial photos in the southern part of the Edwards Plateau (Wermund, and others, 1978). The outcrop in Figure 22 is typical for the Edwards Plateau; it shows the interconnectivity that exists between fractures and bedding planes which gives the Edwards an effective porosity of approximately one percent (Robert Mace; pers. comm., 1995). Secondary porosity forms along fractures and bedding planes, and develops in some horizontal beds by dissolution of burrow fillings and by karst breccia (Sharp, 1990; Maclay and Small, 1984; Rose, 1972). Primary porosity exists within grainstones in the form of Interflow and diffuse flow, move in an intergranular pores. ground-water approximate stair-step fashion down fractures, then laterally within permeable beds and infiltration of rainwater is favored where fractures along bedding planes, intersect the ground surface. Collapse breccia in the Kirschberg Evaporite Member A collapse breccia zone occurs at the top of the Fort Terrett Formation This breccia formed when resistant beds of limestone collapsed in upon (Figure 23). a dissolution zone previously occupied by gypsum that had accumulated in the Lower Cretaceous Kirschberg Lagoon (Figure 17). After the gypsum was dissolved by circulating ground water, the less soluble overlying beds collapsed into the space vacated by the evaporites to form a breccia with a higher permeability (Peter Rose; This breccia allows water to infiltrate the formation at locations pers. comm., 1995). where it has become exposed by erosion. Locally, however, porosity and be reduced where voids have been filled with caliche and soil. permeability may the Because gypsum is completely dissolved in these zones, the hydrochemistry of Edwards aquifer does not reflect contact with gypsum. 45 Figure 22. Hydraulically connected fractures and horizontal bedding planes in the Edwards Group serve as conduits for flow in the unsaturated and saturated zones. A pocket knife at the bottom center of the photograph serves as scale. (Edwards Kirschberg the of Formation Terrett dissolution Fort to the due beds. of here top overlying the exist at of occuring rainwater collapse of breccia infiltration consequent collapse the the for and of Zone Conduits horizon 23. Figure Group).Evaporite Contact springs in the Edwards Group Formations to In the foreground of Figure 24 is a concrete structure that serves capture intermittent spring flow from the Segovia Formation. The concrete structure fills with spring water, and then the water is piped downhill to a stock tank (Figure 25). The source of this water is a contact spring that discharges water from the foot of a colluvial slopewash deposit upslope of the collection structure (Figure 26). This small spring and numerous others occur where bedding planes and/or fractures The intersect the sloping ground surface that truncates the ground-water flowpaths. total discharge from these springs is impossible to measure due to the small size of the their schedules, and physical springs, übiquity, unpredictable discharge inaccessibly. Instead, an estimate of spring discharge as a percentage of recharge a finite difference flow will be made later by assigning drains to a line of cells within model. Discharge from these springs has declined over the last several decades possibly due to the spread of juniper and mesquite trees which use large amounts of soil moisture The of trees observed in (Figure 8). sparse coverage photographs taken earlier in the century, in conjunction with landowner interview reports of higher and more consistent spring discharge in the past, support the contention that spring time due to these trees. A similar conclusion was discharge has declined over reached by Dugas and Flicks (1994) for the Seco Creek Watershed in Texas. They removed trees from an area in the watershed and measured spring discharge before and after the removal. Spring discharge increased after the trees were removed. 48 Figure 24. The concrete structure in the foreground was designed to capture and pipe spring water to a cattle reservoir. Reduced spring flow to this structure in recent decades, possibly due to the growth of phreatophytes, and the associated increased evapotranspiration, necessitated the drilling of a well. 49 Figure 25. Conceptual diagram showing how the limited volume of water exiting a contact spring in the Segovia Formation (Edwards Group) is captured to fill a livestock reservoir. Figure 26. Spring water exits the Segovia Formation (Edwards Group) onto a dry streambed from beneath a coiluviai slopewash deposit. 51 Aerial photographs of tree growth patterns on the sides of mesas Trees growing in distinct patterns are observed in aerial photographs (Figure 27). This pattern is probably related to springs (ground water discharge) and seeps (interflow discharge) along fractures, bedding planes, and higher permeability beds exposed on the mesa slope faces (Figure 28). An example of a discharge point on the face of a tributary canyon is pictured in Figure 6. Note the staining beneath fractures and bedding planes that has been caused by precipitation of minerals from The heaviest concentrations of trees occur on the southeastern discharging water. sides of the mesas because ground water flows along gently dipping bedding planes in that direction. Evaporation from the ground surface, together with evapotranspiration from these trees, constitutes the largest sink in the overall hydrologic cycle here. Exposure of Lower Cretaceous geological contacts The outcrop in Figure 12 shows exposures of the fractured Fort Terrett Formation at top, the unfractured Basal Nodular Member of the Fort Terrett in the middle, and the calcareous siltstone of the Hensel Sand Formation at the bottom. The Basal Nodular Member a more consistent density and lithology than possesses the overlying Burrowed Member of the Fort Terrett Formation (Rose, 1972). It is the lowest occuring bed of several low permeability beds in the Edwards Group that lower value of vertical causes bedding anisotropy (low K /K ). An even anisotropy zx would exist if vertical fractures were provide conduits for downward not present to percolation. Llano River The Llano River is a tributary of the Little Colorado River (Figure 29). It receives water from two mam tributaries, the North Llano River and the South Llano The River, which together drain an area of 4,789-square km. (1,849-square miles). Llano is a perennial, predominantly effluent river, fed by spring discharge from the area. study the of portion a of photograph aerial altitude Low 27. Figure related is pattern this that believed is It growth. tree of pattern the Note rocks. flat-lying the in springs contact via water ground of discharge the to Figure 28. Conceptual diagram showing how phreatophytes grow in distinct rings around mesas of the Edwards Plateau to exploit contact springs that discharge from horizontal bedding planes. 54 River Colorado the to tributary This River. Llano The 29. Figure system. groundwater the for area discharge primary the as serves Edwards aquifer, and baseflow seepage from the Hensel aquifer. It maintains an average daily flow during the low-flow months of December and January of approximately 233,800 cubic meters (8,256,072 cubic feet) at the Junction, IX stream located area gauge, a short distance upstream of the study (Appendix F). The amount of baseflow seepage added to the Llano River as it flows past the study area is estimated in the mathematical model chapter. Potentiometric surfaces Static water levels measured in hundred water wells were over one completed in both the Hensel and the Edwards aquifers (Figures 30 and 31, and Appendix D). Water table elevations were contoured and two separate potentiometric surfaces were defined (Figure 32). Several water wells, completed within both aquifers, were excluded from the dataset because their water levels represented a composite of the potential from each aquifer. Study objectives required development of maps and datasets for distinct aquifers. water tables were in the one for the Edwards Separate mapped area; River aquifer (under the Plateau), and one for the Hensel aquifer (beneath the Llano the Hensel floodplain). It was not possible to measure the potential in the part of aquifer that occurs beneath the Edwards because there are no wells that penetrate the and from Hensel the entire thickness of Edwards, draw the aquifer only. in of the Hensel had to be with the Hydraulic heads this part aquifer predicted numerical flow model. Measured water table elevations in the Edwards aquifer ranged from 584 to 628 meters (1,918 to 2,060 feet). Water table elevations in the Hensel aquifer ranged from 463 to 529 meters (1,520 to 1,737 feet). Farther under the Plateau, back toward the point occuring directly below the topographic divide, it is possible that the potential of the Hensel aquifer increases to an elevation equal to or exceeding that of the bed at the base of the Edwards, but not as high as heads in the confining Figure 30. Locations and project numbers of water wells used to measure the depths to water for water table elevation calculations. Well specifics are listed in Appendix D. 57 Figure 31. Water table elevations calculated using steel tape These elevations measurements of depth-to-water in wells. were used to draw potentiometric surfaces. 58 Potentiometric contours for the unconfined Edwards Figure 32. Group aquifer and the unconfined portion of the Hensel Sand Formation aquifer. Contours are clashed where inadequate well control exists. 59 overlying Edwards aquifer, it is hypothesized, therefore, that there is leakage downward from the higher potential Edwards aquifer to the lower potential Hensel aquifer through this confining bed. This hypothesis is tested in the mathematical model chapter. The average hydraulic gradient in the Edwards aquifer is calculated to be 0.0045, and the average ground-water gradient in the Hensel aquifer is 0.0055. Saturated thickness averages approximately forty meters (130 feet) in the Edwards aquifer, and thirty meters (100 feet) in the Hensel aquifer. Using estimates of 0.0045 for hydraulic gradient, 3.5E-04 cm/sec (one foot/day) for hydraulic conductivity, and 0.01 for effective porosity, a gross estimate of ground-water = velocity using the Darcy seepage velocity equation (v ki/n ) is (3.5E-04 cm/sec * e = 0.0045)/0.01 1.6E-04 cm/sec. (0.45 feet/day) in the Edwards aquifer. For the Hensel aquifer, velocity is estimated at (1.75E-03 cm/sec * 0.0055)/0.10 or 9.7E-05 cm/sec. (0.28 feet/day). Ground-water flowlines A map of approximate ground-water flowlines is presented in Figure 33. Flowlines were drawn orthogonal to potentiometric surface contours (Figure 32). the flow field in Orthogonal flowlines accurately represent horizontally isotropic media. In this case, the Edwards is an anisotropic aquifer, and the Hensel is an anisotropic and heterogeneous aquifer. Therefore, the flowlines are approximations. Flowlines Edwards aquifer converge the steep tributary valleys in the at incised in the Edwards. Flowlines in the Hensel converge at the Llano River. Plotted on the map are the approximate locations for several strong springs that discharge from Edwards aquifer. the Note that these occur where there is a dense of flowlines. convergence 60 Figure 33. Generalized ground-water flowlines within the Edwards Group aquifer and the Hensel Sand Formation aquifer. Flowlines diverge from recharge areas and converge toward discharge areas such as the three mapped springs. 61 Water chemistry analyses Twenty-two wells were selected (Figure 34) for analyses of major ions (Ca, Mg, Na, K, Cl, and S0 ), unstable parameters (pH, Eh, dissolved 0 temperature, 4 2, and alkalinity as HC03), and additional constituents related to REDOX state, flowpath interpretation, and anthropogenic influences (Fe N0 NH, IDS, hardness, 2, 3, and total sulfides). Wells tested included seven wells in the Edwards aquifer, thirteen welts in the Hensel aquifer, and one well in the shallow alluvial aquifer next to the Llano River. Water chemistry analyses were performed to test for evolution of ion concentrations along inferred pathways of ground-water flow, to determine if separate aquifers existed based on the measurement of distinct hydrochemical facies, and to support the mapping of recharge and discharge areas. Field and laboratory procedures used to collect and analyze water samples are listed in Appendix A. The table of analytical values is presented in Appendix G. The error in charge balance between anions and cations for each sample ranged from 0.7 to 5.2 %. A charge balance of 5.0 % percent is considered an acceptable margin of error so the data was judged sufficiently accurate. One of the important purposes of chemistry analysis in this study water was to detect evolution of major ion and unstable parameter concentrations in wells believed to be located along lines parallel to ground-water flowpaths (Back, 1966; Chebotarev, 1955). Chebotarev explained that ground water tends to evolve He observed that this evolution was chemically toward the composition of seawater. normally accompanied by the following regional change in dominant anion species: travel along flowpath (increasing age) . —. HCO~—*HCOSO]~—. SOj~+ HCO ~ —.so;"+ Cl~—*Cl~+ S0 2 CT 4 Figure 34. Locations of wells used to sample ground water for major ions, selected minor ions, and unstable constituents. 63 Wells located along three transects extending from the upland area to the center of the floodplain were included in the analysis (Figure 34). These transects were: 1) wells 1-2-3-(inaccessible zone)-15-16-18, 2 wells 4-7-10-14-19, wells S-6-8-9­11-12-13. it was determined that the hypothesis of ion evolution could be tested here in a limited way only because the flowpaths were too short, and because the abrupt change in hydrochemical facies occuring at the transition from the Edwards to the Hensel wells hides the more subtle changes that could be attributed to anion evolution. A marked contrast in water chemistry between the two aquifers is depicted graphically in the Piper diagram which displays the relative of major concentrations ions (Figure 35, Piper, 1953). Two distinct hydrochemical facies are apparent; the facies of the Edwards aquifer, and the mixed facies of the Hensel Ca-Mg-HC03 aquifer. Some evolution apparently occurs within the Hensel; this can be seen in the cation the the vector Hensel triangle of Piper diagram. Along representing the increases in Na concentration from the edge of the floodplain to the aquifer, water center of the floodplain. This is probably caused by cation exchange of Na ions for silt at Ca ions on smectite clays. Samples of the fine-grained calcareous the top of the Hensel were x-rayed to test this hypothesis (Figure 36). It was determined that does exist smectite, with its high cation exchange capacity, in these samples (Leo evolution occurs as a result of this Lynch; pers. comm., 1995), thus, probably to is observed that not only is there a high Na phenomenon. Referring Figure 37 it concentration in the Hensel aquifer Edwards, high Cl relative to the but also a concentration. High Cl is most likely caused by the dissolution of halite. Likewise, the halite dissolution. The high measured Na concentrations are probably due in part to Hensel semi-arid to arid, near-shore environment where sahbka was deposited in a conditions existed periodically (Stricklin, and others, 1971). It is probable that Na, Mg, Cl, and S0 in the Hensel aquifer (Figures 37 and 38) are derived from the 4 64 Figure 35. Piper diagram illustrating the relative concentrations of major ions measured in ground water from twenty-one wells in the Edwards aquifer and the Hensel aquifer. 65 Figure 36. X-ray diffraction patterns for samples of red siltstone (top) and siltstone (bottom) from an outcrop green at the top of the Hensel Sand Formation. 66 Figure 37. Concentration ranges and median values of major ions in the Edwards and Hensel aquifers. Hensel water samples have higher IDS. 67 Figure 38. Geology map with Stiff diagrams that represent the spatial distribution of absolute ion values and IDS. The higher the IDS, the larger the size of the diagram. 68 dissolution of disseminated veins of evaporitic minerals within low permeability, fine-grained, overbank deposits that have withstood flushing. Elevated concentrations of the ions Mg, Ca, and HC0 exist in both aquifers as a result of dissolution of 3 carbonate rocks by ground water that increases in acidity as it passes through the soil zone and dissolves C0 Drillers’ indicate that thin . 2 logs and geologic exposures layers of carbonate and disseminated caliche exist in the Hensel and yield Mg, Ca, and HC0 ions to the ground water dissolution. The upon comparatively higher 3 concentrations of HC0 3 in the Hensel wells could be caused by bacterially-mediated reduction of sulfate which yields hydrogen sulfide gas and HC03 Hydrogen sulfide . gas occurs commonly in Hensel wells and can be detected by smell at extremely low concentrations. The distribution of electrochemical and dissolved (Eh) potentials oxygen concentrations are The both mapped in Figure 39. relatively high magnitude of parameters in Edwards wells, coupled with the highest hydraulic heads in the study area, demonstrate that this is a zone of recharge. The ground water is probably younger and the aquifer more open to the atmosphere in the Edwards. The Hensel waters reflect reducing conditions; both dissolved oxygen and Eh values are lower. a This water is probably older and the TDS higher as result of longer ground-water residence times (Figure 38). Ground-water samples in Hensel wells have an iron result of the reduction of iron hydroxide and taste and a hydrogen sulfide smell as a sulfate by bacteria in the presence of organic matter. An exception to this overall pattern is observed at the northeastern end of the floodplain (Figures 38 and 39). Here, the dissolved oxygen is intermediate and the Eh is contours exist here, but potential high. Closely-spaced potentiometric that without the higher TDS that might be expected in water is moving through less As it turns Hensel thins in this area as it gets closer to permeable media. out, the Figure 39. Spatial distribution of dissolved oxygen and Eh. High values of both parameters indicate an oxidizing environment; low values, a reducing environment. 70 the Llano Uplift, and there are beds of carbonate rock (recorded in drillers’ logs) that may serve to slow ground-water flow. At the southwestern end of the floodplain the Hensel is thicker, there is more rock (recorded in drillers’ fine-grained sedimentary logs) with disseminated minerals available for dissolution, more coarse-grained and fluvial channels to support rapid throughput of ground water. Analytical results are evolved most easily explained by assuming that ground-water chemistry by cross-bed flow that led to mixing (Figure 40). Figure 40. A contrast in potentiometric surface contour densities, and anomalous TDS concentrations, occur in the ground waters from one end of the Llano River floodplain to the other due to a change in lithologic facies. Pictured is the mechanism believed responsible for the co-occurence of high K and high TDS at the southwest end of the floodplain. 72 Conceptual Model of Ground-Water Flow Physical hydrogeology is Figure 41 a conceptual hydrogeologic cross-section that illustrates the architecture, and the inputs (red), throughputs (green), and outputs (blue) of the ground-water flow system. It is not determinable whether the Edwards aquifer is partially or fully perched because there are no water level data for the portion of the Hensel aquifer that underlies the plateau. Numerical simulations are run to attempt to answer this question. A revised conceptual cross-section of flow based on numerical simulation results is presented in the mathematical model chapter. Rainfall infiltrates portions of the exposed surfaces of the Edwards and Hensel to recharge the aquifers within these formations. The amount of rainfall at recharging the aquifers is calculated approximately 5.06 percent or 3.10 cm/year of mean annual B and The (1.22 inches/year) precipitation (Appendices F). overland remaining rainfall runs off, evaporates, or transpires. Runoff occurs as flow and as channel flow within ephemeral tributaries that meander toward the Llano River. and from The largest portion of the rainfall evaporates from the soil surface trees and plants as they transpire. Rainwater that infiltrates the surface of the Edwards downward in moves the unsaturated zone via vertical fractures and zones of collapse breccia, and then horizontally along bedding planes and permeable beds, until being intercepted by other vertical fractures. Movement therefore, occurs generally in a stair step pattern in the downgradient direction. Where structural or stratigraphic pathways intersect the water exits If escarpment surface in tributary valleys, there in the form of seeps. rainwater infiltrates the flat top of the plateau surface close to the edge of a mesa (thus having a short flowpath) it may be short-circuited (as interflow) to one of these seepage points. The infiltrating rainwater that does not seep out of the 73 Conceptual cross-section of the study area Figure 41. illustrating the architecture, and inputs (red), throughputs (green), and outputs (blue) of the ground-water flow system. 74 system at these points continues to move downward until it recharges the unconfined aquifer. Rainwater that infiltrates the surface at the topographic divides between subwatersheds the It is that these probably reaches aquifer as recharge. likely summits serve as areas of focused recharge to the aquifer because beneath the divides the distance infiltrating water has to travel to reach the underlying saturated zone is less than the distance it would have to travel to arrive at the face of the erosional escarpment to be discharged as seeps. This hypothesis of focused recharge is verified using the numerical model of flow. Results are shown in the mathematical model chapter. Water that does reach the saturated zone of the Edwards probably moves as laminar flow from the carbonate upland toward the Llano River which meanders along the axis of a wide floodplain. There are no large springs in the area which would existence of conduits which would suggest the large subsurface support turbulent flow. Moving laterally along horizontal bedding planes, or downward within vertical fractures, some ground water discharges at the face of the escarpment in contact or fracture springs that occur atop low permeability beds. Some ground water is discharged by phreatophytes, and the remaining ground water recharges the underlying Hensei aquifer by cross-formational flow. Spring discharge and cross­ formational flow are quantified in the numerical model. Downward leakage of ground water to the Hensei aquifer occurs through a fifteen foot thick marly limestone bed (Basal Nodular Member) which is present The Hensei Sand Formation throughout the region forming the base of the Edwards. forms a sediment wedge that is approximately 20 meters (66 feet) thick below the and increases to more than 300 main west-east topographic divide in the study area, meters (1000 feet) in thickness southeast of the Plateau region (Cartwright, 1932). As the saturated thickness of the Hensei aquifer decreases toward the a result, northwest until the formation pinches out beyond the study area. 75 It was not possible to estimate precisely the saturated thickness of the Edwards because wells do not the saturated and aquifer fully penetrate zone, because the thickness of the host formations is variable. A of 20 to 64 meters range (66 to 210 feet) was calculated by subtracting the average elevation of the flat-lying Edwards-Hensel contact from water-table elevations measured in the Edwards aquifer wells. Ground-water flow in the Hensel aquifer at the western end of the floodplain occurs by preferential movement along relatively high permeability sand and gravel fluvial channel deposits encased within lower overbank and sahbka permeability deposits (John Ashworth, pers. comm., 1994). Ground water exits the Hensel aquifer by seeping into the Llano River which operates as the regional base level across which no ground water can flow. A small part of its flow is captured by low livestock and domestic wells which occur at a density of less than capacity one/square mile. It is not a volume of the Hensel water moves likely that great ground downward into the Paleozoic rocks because the Paleozoics are less permeable than the Cretaceous rocks; and they host aquifers which are highly compartmentalized (Barker and Ardis, 1992). Gamma-neutron logs from oil tests fifty miles to the The cross-section of these southwest provide independent evidence for this claim. below the Cretaceous below the Wichita logs demonstrated that (i.e., Paleoplain most of the 150 meters of Unconformity), the log response over upper (500 feet) Paleozoic rocks surpasses the magnitude established as the “shale line.” Therefore, the permeability of the Paleozoic formations is probably low. Contrasting this, the within the overlying Cretaceous rocks is magnitude of the gamma-neutron response uniformly lower than the “shale line” value (Erica Boghici, pers. comm., 1995). The saturated thickness of the unconfined portion of the Hensel aquifer could not be determined exactly. A range of 15 to 70 meters (50 to 225 feet) was 76 calculated by subtracting elevations of the Hensel/Paleozoic contact from elevations of the water table. Chemical hydrogeology Figure 42 is a conceptual cross-section of the study area which shows the dominant hydrochemical species found in each aquifer and the hydrochemical mechanisms believed responsible for their occurance. in the Edwards, water infiltrating the surface dissolves carbon dioxide in the soil zone to reduce the pH of the water that recharges the aquifer. This ground water then slowly dissolves the limestones and dolomites of the Edwards to yield low concentrations of dissolved Ca, Mg, HC0 and total dissolved solids (TDS, 279 to 3, 418 mg/l). High dissolved oxygen (DO) (6.0 to 8.6%) and high electrochemical potential (Eh) (328 to 488 mV), in conjunction with the highest hydraulic heads in the study area, demonstrate that this is a recharge zone (see Figures 32 and 39). In the Hensel aquifer, low concentrations of dissolved Ca, Mg, and HC0 are 3 attributed to the dissolution of the carbonate beds and caliche that can be seen in and cross-formational transfer of Edwards waters. Elevated outcrops, by concentrations of Cl are derived from dissolution of halite that accumulated in nearshore Cretaceous sahbkas (Stricklin and others, 1971). Elevated concentrations of Na are caused by dissolution of halite, and also by ion exchange. Smectite clays are found throughout the Hensel (see diffractogram, Figure 36). These clays possess a high cation exchange capacity, and are probably exchanging bound Na for Ca and Mg in solution. Dissolution of plagioclase feldspar is a possible source of smaller amounts of Na. High concentrations of SC are caused by dissolution of 4 the absence of associated high concentrations of Ca is probably due to loss gypsum; by adsorption. 77 42. Schematic diagram of the gross mechanisms believed to control the occurances and relative concentrations of major ions in the Edwards and Hensel aquifers. 78 The Hensel aquifer is more isolated from the oxidizing influence of the atmosphere than the Edwards aquifer. The iron taste and hydrogen sulfide smell observed in many of the wells, coupled with the low measured concentrations of DO to andEh to228 of (1.1 2.1%) (99 mV), support the interpretation a reducing environment. sulfate and being reduced Apparently, iron hydroxide are during the bacterially mediated oxidation matter. These reactions of organic yield hydrogen sulfide, ferric iron, and bicarbonate to the ground water. Elevated concentrations of TDS (404 to 1603 mg/l) are probably caused by contact with disseminated evaporites, and possibly by a longer ground water residence time in the Hensel compared to the Edwards. Measured concentrations of DO, TDS, and Eh may underestimate actual reducing conditions in the aquifer. This is because the annular spaces around wells over the entire length of the are open wells. As a result ground water in reduced zones will mix with and become diluted by water in zones that are more oxidized. the In the Hensel aquifer the distribution of hydrochemical facies is related to architecture and mineralogy of the rocks through which the ground water flows The higher average value of TDS measured in ground water from wells (Figure 38). at the west end of due to water contact with the floodplain is probably ground overbank clays as it flows preferentially through sinuous sand and gravel deposits that are juxtaposed against these clays (Figure 40). This interpretation is supported by lithologic log data which indicates the existence of these deposits. It is supported also by potentiometric contours which are widely spaced. This wide spacing sands and indicates that flow is occuring through the more highly permeable fluvial gravels recorded in drillers’ logs. portion of the aquifer at end of the The the east with a lower TDS, would be expected to have floodplain, where wells yield water lower density potentiometric high hydraulic conductivity. contours indicative of Drillers’ However, unexpectedly, the contours are tightly spaced (Figure 40). logs 79 show that there is a lithologic facies change to sand and silt with interbedded carbonates, thus there is a west to east transition to a more marine paleoenvironment. During calibration of the numerical model, the process of estimating the spatial distribution of K in the Hensel to produce a potentiometric contour match is very important. 80 Mathematical Model of Ground-Water Flow Introduction For this study a deterministic numerical model used to simulate ground- was water flow in the Edwards and Hensel aquifers. The model simulates flow in a regional system utilizing a set of fundamental approximations. First, the porous medium is assumed to be a continuum that real of replaces the complex system solids and voids. Second, ground-water flow is assumed horizontal because the ratio of aquifer thickness to horizontal length is small. Third, Darcy’s Law is assumed. This equation is: - + + w=0 o>/285 <1347 - - 26 16-Mar w 1665 45 1621 75 1590 - - 27 16-Mar w 1675 45 43 1632 100 1620 - 28 16-Mar w 1735 94 86 1649 100 1635 124 Table of water levels in wells Windmill Bottom Project well # Date measured (1994) (w) Submersible pump (S) Ground elevation (feet asl) Depth to water (state data) (feet) Depth to water (measured) (feet) Water table elevation (feet asl) Total depth (state data) (feet) Total depth (measured) (feet) of hole elevation (calculated) (feet asl) 29 16-Mar w 1740 124 1616 140 1600 - - 30 16-Mar w 1700 90 1610 170 1600 - - 31 16-Mar w 1688 72 68 1620 80 1598 - 32 16-Mar w 1716 105 104 1612 115 1601 - 33 16-Mar s 1770 146 148 1622 200 155 1615 34 17-Mar s 1655 42 1613 120 1535 - - 35 17-Mar s 1840 103 1737 150 113 1727 - 36 17-Mar w 1913 222 1691 450 280 1633 - 37 17-Mar w 1840 110 104 1736 127 130 1710 38 17-Mar w 2056 100 1957 180 1876 - - 39 17-Mar s 1645 31 1614 80 1565 - - 40 17-Mar s 1666 39 1627 90 1576 - - 41 18-Mar s 1650 60 1590 110 1540 - - 42 19-Mar w 2018 68 1950 120 1898 - - 43 19-Mar w 1910 15 1895 60 1850 - - 44 19-Mar s 1845 208 1637 250 1595 - - 45 19-Mar w 2020 370 1650 >400 1620 - - 46 20-Mar w 2052 120 1932 168 165 1887 - 47 5-Apr s 1704 - 100 1604 - - - 48 5-Apr s 1640 - 41 1599 - - - 49 5-Apr s 1660 - 54 1606 - - - 50 6-Apr s 1680 68 58 1622 101 - 1579 51 6-Apr s 1695 - 73 1622 - - - 52 6-Apr s 1680 - 61 1619 - - - 53 7-Apr s 1680 - 62 1618 - - - 54 7-Apr s 1700 - 83 1617 - - - 55 7-Apr s 2010 - 92 1918 - - - 56 8-Apr s 1670 27 1643 100 1570 125 Table of water levels in wells Project well # Date measured (1994) Windmill (w) Submersible pump (s) Ground elevation (feet asl) Depth to water (state data) (feet) Depth to water (measured) (feet) Water table elevation (feet asl) Total depth (state data) (feet) Total depth (measured) (feet) Bottom of hole elevation (calculated) (feet asl) 57 8-Apr w 1640 57 1583 90 1550 - - 58 8-Apr s 1642 55 54 1588 105 72 1570 59 8-Apr s 2050 198 1852 450 400 1650 - 60 25-Apr s 2170 188 1982 239 1931 - - 61 25-Apr s 1996 - 174 1966 - 200 1796 62 25-Apr w 1670 73 51 1619 110 - 1560 63 25-Apr w 110 190 - - - - - 64 25-Apr w 1990 10 8 1982 20 20 1970 65 26-Apr s 1760 - 43 1717 54 - 1706 66 26-Apr s 1760 32 1728 200 1560 - - 67 26-Apr w - 55 60 - 100 - - 68 26-Apr s 2083 100 132 1951 210 200 1883 69 1-May w 2320 - 260 2060 340 - 1980 70 1-May s 2225 140 217 2008 220 270 1955 71 1-May w 2150 149 146 2004 200 235 1915 72 1-May s - - 34 - - 80 - 73 1-May w 2230 - 228 2002 380 285 1850 74 1-May w 2151 - 126 2025 - 150 2001 75 14-May w 2160 191 1969 350 - 1810 76 14-May w 2064 - 121 1943 200 220 1844 77 14-May w 2185 - 185 2000 - 225 1960 78 15-May s 1720 - 96 1624 - 130 1590 79 15-May w 1720 - 93 1627 - 110 1610 80 15-May s 1782 - 161 1621 - 190 1592 81 17-May w 2188 - 156 2032 170 240 1948 82 17-May w 2185 - 187 1998 - 260 1925 83 17-May s 2310 - 266 2044 - 275 2035 84 17-May s 2071 101 93 1978 200 204 1867 126 Table of water levels in wells Windmill Bottom Project well # .ate measured (1994) (w) Submersible pump (s) Ground elevation (feet asl) Depth to water (state data) (feet) Depth to water (measured) (feet) Water table elevation (feet asl) Total depth (state data) (feet) Total depth (measured) (feet) of hole elevation (calculated) (feet asl) 85 17-May s 2088 225 122 1966 125 206 1882 86 17-May s 2010? 88 1922? 250 1760 - - 87 17-May w 2080 102 97 1983 225 166 1914 88 17-May w 2194 240 - 1954 225 275 1919 89 17-May w 2190 - 184 2006 - 205 1985 90 state data s 2140 155 1985 240 1900 - - 91 state data w 1797 183 1614 240 1554 - - 92 state data s 1666 66 1600 118 1548 - - 93 state data s 1720 96 1624 150 1570 - - 94 state data s 1600 46 1554 120 1480 - - 95 18-Jul w 2270 280 1990 320 1950 - - 96 18-Jut w 2190 178 2012 195 1995 - - 97 18-Jul w 2209 205 146 2063 220 225 1984 98 18-Jul w 2165 195 1970 254 1911 - - 99 18-Jul w 2135 142 144 1991 200 1935 - 100 18-Jul w 2064 50 2014 200 1864 - - 101 18-Jul w 2070 149 1921 220 1850 - - 102 state data w 2253 238 2015 250 2003 - - 103 state data w 2250 256 1996 >280 <1970 - - 104 state data s 2190 186 2004 264 1926 - - 105 state data s 1680 38 1642 104 1576 - - 106 state data s 1665 40 - 1625 65 - 1600 107 statedata s 1640 42 - 1598 60 - 1580 108 state data s 1620 40 - 1580 60 - 1560 109 state data s 1617 40 - 1577 50 - 1567 110 state data s 1646 88 - 1558 220 - 1426 111 state data s 1580 60 - 1520 324 - 1256 112 state data w 2092 126 1966 240 1852 127 Appendix E Specific capacity and transmissivity calculations 128 Specific capacity/ transmissivity calculations (based on driller sc test values used within the Walton (1962) equation) Edwards Group Formations Transmissivity (feetA 2/day) assuming the following values of storativity: Specific capacity= rate of Well # Pumptime (days) Well radius (leet) discharge(Q) in gpm/ S = .02 S = .002 S = .0002 S = .00002 drawdown(s) in feet 56-27-1 a 0.0139 0.29 0.0340 1.9 3.3 4.7 6.0 56-27-2 0.0208 0.21 0.1000 9.6 13.6 17.5 21,4 56-19-4a 0.0417 0.25 0.2500 35.5 39.4 49.1 58.6 56-19-4a 0.0833 0.25 0,0667 7.2 9.9 12.6 15.1 56-19-7b 0.0208 0.25 0.1692 16.7 23.5 30.0 36.6 56-19-7b 0.0208 0.25 0.0465 3.5 5.5 7.3 9.2 56-19-7a 0.0208 0.25 0.0543 4.1 6.4 8.5 9.6 56-19-5a 0.0208 0.25 0.1300 12.6 17.9 23.2 28.4 56-19-5b 0.0417 0.25 0.0170 1.2 1.9 2.5 3.2 56-19-71 0.0313 0.25 0.0400 3.6 5.5 7.2 8.9 56-19-71 0.0208 0.25 0.1200 10.9 15.9 20.7 25.2 56-19-7e 0.0208 0.25 0.0100 0.4 0.9 1.3 1.7 56-18-6c 0.0313 0.25 0.1000 9.6 13.6 17.6 21.5 56-18-6a 0.0833 0,25 0.1091 12.6 17.0 21.1 25.2 56-18-6a 0.0417 0.25 0.3000 35.0 46.7 58.1 70.1 56-18-9e 0.0208 0.30 0.0667 6.3 8.9 11.7 14.2 56-18-9c 0.0208 0.20 0.0042 0.3 0.4 0.7 0.9 56-18-9c 0.0104 0.25 0.0057 0.1 0.4 0.7 0.9 56-18-9c 0.0104 0.25 0.0042 0.1 0.4 0.5 0.7 56-18-9c 0.0104 0.25 0.0083 0.1 0.1 0.1 0.1 56-18-9b 0.0417 0.25 0.1406 15.1 20.8 26.2 31.5 56-26-3a 0.0417 0.25 0.0571 5.2 7,5 9,7 12.0 56-19-8a 0.0208 0.25 0.1100 9.9 14,4 18.8 23.0 56-19-8g 0.0208 0.25 0.0300 2.0 3.2 4.4 5.6 56-19-8d 0.0208 0.25 0.2300 23.9 33.2 42.2 50.9 56-19-8c 0.0208 0.25 0.0500 3.7 5.9 7.9 9.7 sums: 231.1 316,3 404.5 490.4 averages: 8.9 12.2 15.6 18.9 129 Specific capacity/ transmissivity calculations Hensel Sand Formation Transmissivity (feetA 2/day) assuming the following values of storativity: Specific capacity= Well# Pump time (days) Well radius (feet) rate of discharge(Q) in gpm/ S = .02 S = .002 S = .0002 S = .00002 drawdown(s) in feet 56-27-6) .0417 .2080 .4167 56.0 72,4 88.6 104.4 56-27-6e .0208 .2090 .2375 26,6 36.3 45.5 54.8 56-27-6e .0208 .2090 .2500 28.3 38.4 48.2 57.6 56-27-6e .0208 .2090 .2368 26.5 36.1 45.4 54.4 56-27-6C .0208 .2090 .1111 10.9 15.5 19.9 24.1 56-27-6C .0208 .2090 .1333 13.5 19.0 24.1 29.2 56-27-6b .0208 .2920 .0615 4.6 7,3 9.7 12.1 56-27-6f .0104 .2090 .0167 0.8 1.6 2.3 3.0 56-27-6g .0104 .2090 .0822 6.6 10.0 13.2 16.6 56-27-81 .0208 .2090 2.0000 298.8 376.9 453.6 530.1 56-28-1 d .0208 .2290 3.3333 516.7 646.8 773.8 900.3 56-28-1 d .0208 .2500 .0909 10.6 11.8 15.3 18.8 56-28-1 f .0208 .2290 2.8571 435.4 546.8 656.1 764.7 56-28-1 g .0417 .2920 5.8333 980.5 1206.6 1428.7 1648.6 56-28-1 h .0833 .2090 5.0000 942.8 1134.6 1324.2 1512.1 56-28-1 h .0417 .2090 10.0000 1886.3 2269.9 2648.6 3024.9 56-28-1 k .0208 .2090 .4000 48.7 64.5 80,1 95.4 56-28-11 .0208 .2500 1.3330 181.6 233.9 285.2 376.4 56-28-1 m .0104 .2500 .0833 6.1 9.6 12.9 16.2 56-28-Ip .0208 .2500 .1000 9.0 13.2 17.0 23.1 56-28-2b .0139 .2770 .0781 5.7 9.0 12.1 15.2 56-28-2b .0139 .2770 .1786 16.0 23.4 30.4 37.4 56-28-4 .0417 .2090 .0247 2.0 3,1 4.0 5.1 56-28-4f .0208 .2090 .0147 0.9 1.6 2.2 2.7 56-28-4f .0104 .2090 .0222 1.2 2.2 3.1 3.9 56-28-4h .0208 .2090 .0400 3.2 4.8 6.5 7.9 56-28-4k .0208 .2090 .0455 3.6 5.5 7.3 9.0 56-28-4p .0417 .2090 .0054 0.3 0.5 0,7 0.9 56-27-9C .0208 ,2500 .0079 0.3 0.7 0.9 1.3 56-27-9a .0208 .2500 1.2000 161.0 208.3 254.8 300.5 56-28-4b .0417 .2500 .0960 9.7 13.7 17.4 21.1 56-28-4C .0417 .2500 .1600 24.1 30.4 36.5 42.5 56-28-4 .0208 .2500 .0532 4.2 6.3 8.5 10.6 130 56-28-7a .0417 .2500 .8000 111.3 142.8 173.6 204.0 56-28-7C .0208 .2500 3.0000 450.1 567.3 682.5 796.2 56-28-4q .0208 .2090 .2778 31.9 43.1 53.8 64,6 56-28-4r .0208 .2090 .1429 14.7 20.6 26.1 31.6 56-28-4S .0208 .2090 .0079 0.4 0.8 1.1 1.3 56-28-4 .0208 .2090 .0532 4.4 6.7 8.6 10.8 56-20-6b .0417 .2500 .5000 65,5 85.3 104.7 123.7 56-20-7a .0208 .2500 .0800 6.5 9.7 12.8 15.5 56-20-8 .0104 .2500 .0900 7.0 10,9 14.7 18.3 56-20-8 .0069 .2500 3.1300 412.0 535.7 656.7 775.8 56-20-8 .0104 .2500 1,0000 119.0 158.8 197.8 236.1 56-20-4b .0417 .2500 .8500 119.1 152.6 185.3 217.6 56-20-51 .0625 .2500 .8600 126.8 160.4 193.5 226.1 56-20-5q .0208 .2500 ,1700 16,8 23.8 30.4 37.0 56-20-5p .0104 .2500 .0200 0.8 1.7 2.6 3.4 56-20-5m .0208 .2500 .0500 3.8 5.9 7.9 9.8 56-20-5h .0208 .2500 .0900 7.8 11.6 15.2 18.7 56-20-5g .0417 .2500 .7500 91.8 121.7 150.9 179.5 56-20-5g .0208 .2500 .1900 19.2 27.1 34.5 41.7 56-20-5e .0208 .2500 .3400 38.0 51.7 65.0 78.1 56-20-5e .0417 .2500 .7000 95.8 123.4 150.3 177.1 56-20-5d .0417 .2500 1.3330 196.8 248.9 300.0 350.3 56-20-5C .0417 .2500 1.0000 143.1 182.4 220.9 258.8 56-20-5a .0417 .2500 1.3300 196.8 248.9 300.0 350.5 sums: 8001.6 10002.4 11965.4 13952.0 averages: 140.4 175.5 209.9 244.8 131 Appendix F calculations Aquifer recharge 132 Calculation of average monthly rainfall and average monthly Llano River discharge during January and December (low-flow months) for the period 1939-1988 for the combined North and South Llano River watersheds west of the study area. Sources of data: Rainfall Dunk family for the National Weather Service - USGS river # 08150000 - River discharge gauge High discharge values caused by storm runoff are removed and replaced by the ave. of Llano River the remaining monthly Month Year Rainfall (inches) Discharge (acre-feet' month) discharges. This adjusts the database in a way that allows the ave. river discharge to be more representative of gw discharge only. JAN 1939 2.51 7210 DEC 1939 2.29 6600 JAN 1940 0.58 6280 DEC 1940 2.27 5500 JAN 1941 0.45 4820 DEC 1941 0.78 5610 JAN 1942 0.00 5270 DEC 1942 0.85 7360 JAN 1943 0.32 6650 DEC 1943 5330 1.90 JAN 7860 1944 4.10 DEC 1944 1.55 4830 JAN 1945 1.78 4920 DEC 1945 0.36 2890 1.91 3640 JAN 1946 0.41 3590DEC 1946 4.18 6060JAN 1947 1.60 3560DEC 1947 0.98 3490JAN 1948 4800DEC 1948 0.07 4840JAN 1949 3.46 6380DEC 1949 2.23 JAN 1950 0.62 6310 DEC 1950 3850 0.00 3730 0.00 1951JAN 0.33 2380DEC 1951 2370 0.09 1952 JAN 1952 3.02 2280 JAN 1953 0.00 2200 DEC 1953 DEC 0.30 2040 0.46 2050 JAN 1954 0.00 1920DEC 1954 1990 1955 1.43 JAN 29801955 0.41 DEC 0.31 2850 1956 JAN 0.80 1560 1956DEC 1610 0.65 1957JAN 5800 11140 0.47 1957 DEC 8550 2.96 1958 JAN 133 DEC 1958 0.69 5800 10780 JAN 1959 1.43 9460 DEC 1959 2.34 6570 JAN 1960 1.29 7390 DEC 1960 4.03 5800 14490 JAN 1961 2.44 5800 13360 DEC 1961 0.55 8730 JAN 1962 0.18 7740 DEC 1962 0.63 4810 JAN 1963 0.00 5120 DEC 1963 0.43 4430 JAN 1964 1.30 4220 DEC 1964 0.55 7320 JAN 1965 2.29 6670 DEC 1965 1.28 4770 JAN 1966 0.87 4670 DEC 1966 0.00 5710 JAN 1967 0.25 5770 DEC 1967 1.68 4470 JAN 1968 6.05 5800 39390 DEC 1968 0.00 5540 JAN 1969 0.25 5540 DEC 1969 1.81 5800 12900 JAN 1970 0.39 5800 11810 DEC 1970 0.00 6650 JAN 1971 0.00 6770 DEC 1971 1.12 9650 JAN 1972 0.77 8780 DEC 1972 0.00 7500 JAN 1973 2.27 7410 DEC 1973 0.00 9200 JAN 1974 0.60 8540 DEC 1974 1.55 5800 16460 JAN 1975 0.00 5800 13490 DEC 1975 0.41 5800 11310 JAN 1976 0.46 5800 10810 DEC 1976 1.59 5800 12530 JAN 1977 1.17 5800 11300 DEC 1977 0.00 9370 JAN 1978 0.47 9210 DEC 1978 1.06 7640 JAN 1979 0.71 7330 DEC 1979 1.76 6290 JAN 1980 1.64 6330 DEC 1980 1.23 7350 JAN 1981 0.93 6920 DEC 1981 0.00 5800 10650 JAN 1982 0.00 5800 10070 DEC 1982 1.47 6880 JAN 1983 0.82 6840 DEC 1983 2.02 5600 JAN 1984 1.18 5440 DEC 1984 7.48 5800 75560 JAN 1985 1.02 5800 20140 DEC 1985 0.00 6770 JAN 1986 0.60 6360 DEC 1986 3.24 5800 14190 JAN 1987 0.00 5800 16580 DEC 1987 0.89 5800 12260 JAN 1988 0.14 5800 11470 DEC 1988 0.54 8880 averages = 2.9 cm (1.14 in.)/ 5686 acre-ft/ month month 134 Estimation of annual average ground-water recharge to be used as the initial value for numerical model input Calculaterecharge inthedrainagebasin above USGS Llano River # 08150000 gauge (drainage basin next to study area which includes drainage from both the North and South Llano Rivers) area of drainage basin above 1,849 square miles or Llano River stream gauge 1,183,360 acres average monthly rainfall during the low-flow 1.14 inches or .0950 feet months of Jan. and Dec. (1939-1988) average monthly rainfall in acre-feet .0950 feet * 1,183,360 acres = 112,419 acre-feet average monthly discharge at river gauge 5,686 acre-feet (Jan and Dec, 1939-1988). Anomalously high discharge values which are caused by surface water runoff are replaced by the average of the remaining stream gauge values % rainfall measured at Llano River gauge 5,686 acre-feet / 112,419 acre-feet = 5.06 % 5.06 % IS THE PERCENTAGE OF RAINFALL DISCHARGING FROM THE AQUIFER INTO THE LLANO RIVER. BECAUSE THIS AQUIFER IS BEING MODELED AS A STEADY-STATE SYSTEM, DISCHARGE = RECHARGE. Now transfer this discharge value to the area under study (which has no river gaugedowngradient)tobeusedastheinitialestimateofrecharge in each cell of the numerical model. annual rainfall in study area 24.15 inches average 1939-1988 (Dunk data for the NWS) 24.15 inches ’ .0506 average annual recharge in study area =1.22 inches or 3.10 cm 135 Appendix G Table of water chemistry values 136 Table of water chemistry values Eh Eh of Project well # Formation PH H + (mg/l) Eh measured (mv) calculated approxi­mation (mv) Zobell standard (mv) Temp (°C) 02(probe) (mg/l) 02 (mg/l) 1 Edwards 7.34 4.57E-08 315 488 255 21.4 8.6 7.0 Limestone 2 Edwards 7.29 5.13E-08 248 414 262 20.3 7.0 7.0 Limestone Edwards 3 7.45 3.55E-08 282 431 279 21.7 8.0 7.0 Limestone Edwards 4 Limestone 7.29 5.13E-08 195 328 295 22.3 6.3 7.0 Edwards 5 Limestone 7.46 3.47E-08 260 416 272 20.5 6.8 6.0 Edwards 6 7.35 4.47E-08 251 419 260 22.3 2.9 2.0 Limestone Edwards 7 Limestone 7.52 3.02E-08 235 404 259 20.6 6.0 6.5 Hensel 8 7.10 7.94E-08 101 255 275 20.1 1.3 0.4 Sand Hensel 9 Sand 7.17 6.76E-08 235 415 248 21.3 5,6 5.5 Hensel 10 7.30 5.01 E-08 45 213 260 22.1 1.1 0.0 Sand Hensel 11 7.16 6.92E-08 325 413 340 21.4 5.0 9.0 Sand Hensel 12 6.95 1.12E-07 310 461 277 22.5 7.1 Sand - Hensel 13 Sand 7.45 3.55E-08 310 413 345 21.5 1.9 0.8 14 Hensel 7.10 7.94E-08 -75 99 254 21.6 1.7 0.0 Sand 15 Hensel 7.24 5.75E-08 -30 108 290 22.0 1.2 0.5 Sand Hensel 16 7.07 8.51 E-08 -5 151 272 21.6 1.8 0.5 Sand Hensel 17 Sand 7.20 6.31E-08 59 193 294 21.4 2.1 1.0 18 Alluvium 7.23 5.89E-08 195 388 235 20.8 1.9 1.5 19 Hensel Sand 7.05 8.91 E-08 255 360 323 21.1 6.5 5.5 21 Hensel Sand 7.40 3.98E-08 250 228 400 21.6 1.3 0.3 22 Hensel Sand 7.45 3.55E-08 178 353 367 22.3 1.6 0.3 137 Table of water chemistry values Project well * Formation 02 % sat. Fe 2+ (mg/l) Fe-tot (mgyi) NH3 (mg/I) Sulfide (mg/l) SI02 (mg/l) Mn (mg/l) HC03 (mg/l) 1 Edwards Limestone 100 .00 .00 0.21 .03 15 .00 269 Edwards 2 79 .05 .08 0.36 .03 12 .00 319 Limestone Edwards 3 .02 .08 0.14 .02 21 .00 302 Limestone - Edwards 4 76 .14 .02 0.25 .03 12 .00 299 Limestone Edwards 5 Limestone 78 .02 .08 0.03 .03 15 .00 236 Edwards 6 Limestone - .11 .11 0.40 .03 14 .00 328 7 Edwards 70 .08 .05 0.32 .03 10 .00 275 Limestone Hensel 8 15 .14 .14 0.60 03 14 .00 352 Sand Hensel 9 60 .02 .05 0.36 .04 15 .00 350 Sand Hensel 10 14 .66 .86 1.25 .04 9 .00 338 Sand Hensel 11 Sand - .05 .11 0.48 .04 12 .00 352 Hensel 12 82 .08 .11 0.40 .03 18 .00 446 Sand 13 Hensel 22 .05 .14 0.73 .04 6 .00 345 Sand 14 Hensel 20 .51 1.85 1.74 .06 0 .00 386 Sand 15 Hensel 14 .05 .74 1.35 .03 9 .00 331 Sand 16 Hensel Sand 21 .11 .24 1.46 .04 9 .00 345 17 Hensel 25 .05 .41 1.46 03 9 .00 327 Sand 18 Alluvium 23 .05 .14 0.25 .06 12 .00 266 19 Hensel 77 .14 .27 0.17 .03 20 .00 333 Sand 21 Hensel .05 .11 1.70 .03 12 .00 303 Sand - 22 Hensel Sand _ .02 .02 1.30 .03 10 .00 368 138 Table of water chemistry Table of water chemistry Table of water chemistry values Project well # Formation HC03 (meq/1) (mm/l) F (mg/I) F (meq/l) (mm/l) Cl (mg/I) Cl (meq/1) (mm/l) N02 (mg/I) Br (mg/I) N03 (mg/1) i Edwards 4.41 0.49 .03 19.74 0.56 0.21 0.20 1.59 Limestone Edwards 2 5.23 0.72 .04 25.42 0.72 0.20 0.23 11.47 Limestone Edwards 3 Limestone 4.95 0.80 .04 31.93 0.90 0.20 0.22 4.65 Edwards 4 Limestone 4.90 0.40 .02 16.65 0.47 0.00 0.00 2.14 Edwards 5 3.87 0.59 .03 31.83 0.90 0.24 0.03 2.13 Limestone Edwards 6 5.38 1.39 .07 34.74 0.98 0.42 0.01 1.59 Limestone Edwards 7 4.51 Limestone - - - - - - - Hensel 8 Sand 5.77 1.55 .08 61.54 1.73 0,39 0.71 0.62 Hensel 9 Sand 5.74 0.50 .03 46.84 1.32 0.44 0.03 9.35 10 Hensel Sand 5.54 1.78 .09 316.73 8.92 0.85 2.17 0.31 Hensel 11 5.77 0.47 .02 38.84 1.09 0,35 0.01 8.13 Sand Hensel 12 Sand 7.31 0.00 .00 36.30 1.02 0.43 0.00 21.80 13 Hensel 5.66 1.84 .10 151.01 4.25 0.00 1.05 2.35 Sand 14 Hensel 6.33 1.32 .07 156.50 4.41 0.43 0.57 0.26 Sand 15 Hensel Sand 5.43 1.72 .09 196.75 5.54 0.53 0.96 0.25 16 Hensel Sand 5.66 1.37 .07 182.05 5.13 0.52 1.45 0.33 17 Hensel Sand 5.36 3.39 .18 242.17 6.82 0.94 1.34 0.74 18 Alluvium 4.36 0.39 .02 32.65 0.92 0.16 0.29 0.53 Hensel 19 Sand 5.46 - - - - - - - 21 Hensel Sand 4.97 2.93 .15 104.99 2.96 0.52 0.54 0.30 22 Hensel 6.03 3.16 .17 104.74 2.95 0.55 0.52 0.24 Sand 139 values Project well # Formation N03 (meq/t) (mnritl) S04 (mg/1) 504 (meq/l) 504 (mm/l) Li (mg/l) Na (mg/l) Na (meq/l) (mm/t) K (mg/1) 1 Edwards Limestone .026 11.2 0.23 0.12 .017 17.876 0.777 1.520 Edwards 2 .185 12.4 0.26 0.13 .018 20.640 0.897 0.850 Limestone Edwards 3 .075 13.7 0.29 0.14 .027 22.190 0.965 0.840 Limestone 4 Edwards .035 7.5 0.16 0.08 .000 13.730 0.597 1.520 Limestone Edwards 5 Limestone 034 21.5 0.45 0.22 .019 27.910 1.213 1.370 Edwards 6 .026 56.1 1.17 0.58 .063 41.650 1.811 4.210 Limestone Edwards 7 .000 21.640 0.941 1.210 Limestone - - - - Hensel 8 .010 110.4 2.30 1.15 .000 68.980 2.999 7.820 Sand Hensel 9 .151 40.0 0.83 0.42 .035 30.580 1.330 0.700 Sand Hensel 10 .005 109.6 2.28 1.14 .000 170.360 7.407 12.410 Sand Hensel 11 Sand .131 26.9 0,56 0.28 .027 28.750 1.250 1.310 Hensel 12 .352 24.9 0.52 0.26 .023 47.780 2.077 7.590 Sand Hensel 13 .038 63.2 1.32 0.66 .111 160.100 6.961 10.780 Sand Hensel 14 .004 188.0 3.91 1.96 .082 94.260 4.098 9.890 Sand Hensel 15 Sand .004 181.3 3.77 1.89 .000 131.580 5.721 14.710 16 Hensel Sand .005 174.9 3.64 1.82 .082 83.100 3.613 9.880 Hensel 17 .012 625.3 13.02 6.51 .203 361.650 15.724 14.310 Sand 18 Alluvium .009 22.0 0.46 0.23 .028 22.570 0.981 1.660 19 Hensel - - - .015 22.450 0.976 3.680 Sand - 21 Hensel Sand .005 390.1 8.12 4.06 .143 190.840 8.297 16.380 22 Hensel .004 389.3 8.10 4.05 .130 196.200 8.530 18.060 Sand 140 values Project well # Formation K (meq/l) (mm/l) Ca (mg/l) Ca (meq/l) Ca (mm/l) Mg (mg/l) Mg (meq/l) Mg (mm/l) Sr (mg/l) 1 Edwards Limestone .04 48.58 2.429 1.212 30.84 2.538 1.269 0.2 2 Edwards .02 83.40 4.170 2.081 24.28 1.998 0.999 0.3 Limestone 3 Edwards Limestone .02 61.53 3.077 1,535 32.41 2.667 1.333 0.7 4 Edwards .04 59.41 2.971 1.482 31.44 2.588 1.293 0.3 Limestone Edwards 5 Limestone .04 43.53 2.177 1.086 27.23 2.241 1.120 0.4 Edwards 6 Limestone .11 57.09 2.855 1.424 42.23 3.476 1.737 4.0 Edwards 7 Limestone .03 50.69 2.535 1.265 33.86 2.787 1.393 0.9 Hensel 8 Sand .20 62.39 3.120 1.557 47.63 3.920 1.959 5.5 9 Hensel Sand .02 76.80 3.840 1.916 42.20 3.473 1.736 0.5 Hensel 10 .32 91.35 4.568 2.279 72.32 5.952 2.975 10.7 Sand Hensel 11 .03 73.88 3.694 1.843 40.07 3.298 1.648 0.6 Sand 12 Hensel Sand .19 83.90 4.195 2.093 42.08 3.463 1.731 0.8 Hensel 13 Sand .28 44.56 2.228 1.112 24.37 2.006 1.002 2.5 Hensel 14 .25 94.08 4.704 2.347 75.02 6.174 3.086 6.3 Sand 15 Hensel .38 78.30 3.915 1.954 66.22 5.450 2.724 13.9 Sand 16 Hensel Sand .25 93.17 4.659 2.325 77.10 6.346 3.172 10.3 17 Hensel Sand .37 85.41 4.271 2.131 77.56 6.384 3.190 22.8 18 Alluvium .04 65.78 3.289 1.641 23.19 1.909 0.954 1.8 19 Hensel .09 87.86 4.393 2.192 26.91 2.215 1.107 0,9 Sand 21 Hensel .42 72.01 3.601 1.797 57.64 4.744 2.371 17.5 Sand 22 Hensel Sand .46 71.42 3.571 1.782 58.01 4.774 2.386 17.6 141 Table of water chemistry values HC03­ Project well # Formation 2(Ca+Mg­S04+.5(Na­Cl)) Ca+Mg (Ca+Mg­S04+ .5"(Na-CI)) (Ca+Mg­S04+.5(Na­CI))/HC03 Ca+Mg/HC 03 Na/CI Ca/S04 Ca/HC03 1 Edwards Limestone -0.540 2.481 2.475 0.561 0.563 1.398 10.413 0.275 2 Edwards -0.853 3.080 3.041 0.582 0.589 1.253 16.139 0.398 Limestone Edwards 3 Limestone -0.566 2.868 2.758 0.557 0.579 1.073 10.769 0.310 4 Edwards -0.622 2.776 2.762 0.563 0.566 1.273 19.047 0.302 Limestone 5 Edwards -0.413 2.206 2.141 0.553 0.570 1.354 4.853 0.281 Limestone Edwards 6 -0.610 3.162 2.994 0.557 0.588 1.851 2.438 0.265 Limestone Edwards 7 Limestone - 2.658 - - 0.589 - - 0.281 8 Hensel Sand -0.230 3.516 3.000 0.520 0.609 1.730 1,355 0.270 9 Hensel -0.745 3.652 3.241 0.565 0.637 1.008 4.606 0.334 Sand 10 Hensel -1.170 5.254 3.356 0.606 0.948 0.830 1.998 0.411 Sand 11 Hensel -0.808 3.492 3.289 0.570 0.605 1.142 6.579 0,319 Sand 12 Hensel -0.874 3.824 4.093 0.560 0.523 2.032 8.080 0.286 Sand Hensel 13 Sand 0.037 2.114 2.809 0.497 0.374 1.636 1.689 0.197 14 Hensel -0.314 5.433 3.321 0.525 0.859 0.930 1.199 0.371 Sand Hensel 15 -0.334 4.678 2.880 0.531 0.862 1.032 1.035 0.360 Sand 16 Hensel Sand -0.181 5.496 2.918 0.516 0.972 0.705 1.277 0.411 17 Hensel Sand -1.167 5.321 3.264 0.609 0.993 2.305 0.327 0.398 18 Alluvium -0.433 2.595 2.397 0.550 0.595 1.067 7.168 0.376 19 Hensel Sand - 3.299 - - 0.604 - - 0.402 21 Hensel -0.587 4.168 2.777 0.559 0.839 2.806 0.442 0.362 Sand 22 Hensel Sand 0.221 4.168 2.906 0.482 0.691 2.891 0.440 0.295 142 Table of water chemistry values Project well # Formation Sr (meq/l) Sr (mm/l) Hardness (mg/l) %Na SAR RSC IDS (mg/l) Ionic Strength (mm) 1 Edwards .005 .002 249 13.4 0.5 -0.6 279 0.008116 Limestone 2 Edwards .007 .004 309 12.7 0.5 -0.9 348 0.009968 Limestone 3 Edwards .015 .007 288 14.3 0.6 -0.8 338 0.009513 Limestone 4 Edwards Limestone .006 .003 278 9.6 0.4 -0.7 292 0.008744 Edwards 5 Limestone .008 004 222 21.4 0.8 -0.5 287 0,007908 Edwards 6 .090 .045 321 22.0 1.0 -1.0 418 0.011769 Limestone Edwards 7 .019 .010 267 15.0 0.6 -0.8 Limestone - - Hensel 8 .124 .062 359 29.3 1.6 -1.3 553 0.014851 Sand Hensel 9 .011 .005 367 15.4 0.7 -1.6 435 0.012438 Sand Hensel 10 .245 .122 539 40.6 3.2 -5.0 961 0.024178 Sand Hensel 11 Sand .013 .006 351 15.1 0.7 -1.2 404 0.011708 Hensel 12 .018 .009 384 20.9 1.1 -0.3 502 0.013663 Sand Hensel 13 .057 .029 215 60.7 4.8 1.4 636 0.014243 Sand Hensel 14 Sand ,145 .072 552 26.9 1.8 -4.6 815 0.022507 15 Hensel .316 .158 485 37.0 2.6 -3.9 856 0.022026 Sand Hensel 16 .234 .117 562 24.3 1.5 -5.3 811 0.022231 Sand 17 Hensel Sand .521 .261 559 58.8 6.8 -5.3 1603 0.038414 18 Alluvium .041 .020 262 15.8 0.6 -0.8 313 0.008856 19 Hensel Sand .020 .010 332 12.7 0.5 -1.1 - - 21 Hensel Sand .399 .200 438 48.6 4.1 -3.4 1014 0.025256 22 Hensel Sand .403 .201 438 49.2 4.2 -2.3 1050 0.025917 143 Table of water chemistry values Sum of Sum of Charge Project well# Formation Cations Anions Balance Error Na/CI Mg/Ca Ca+Mg-S04 Na-CI .5(Na-CI) (meq) (meq) (%) Edwards 1 Limestone 5.79 5.25 4.9 1.398 1.047 2.364 0.221 0.111 Edwards 2 Limestone 7.09 6.43 4.9 1.253 0,480 2.951 0.181 0.091 Edwards 3 Limestone 6.75 6.25 3.8 1.073 0.868 2.726 0.065 0.033 Edwards 4 6.20 5.58 5.2 1.273 0.873 2.698 0.128 0.064 Limestone Edwards 5 Limestone 5.67 5.28 3.6 1.354 1.031 1.982 0.317 0.158 Edwards 6 Limestone 8.34 7.62 4.5 1.851 1.220 2.577 0.832 0.416 Edwards 7 Limestone 6.31 - - - 1.101 - - - Hensei 8 Sand 10.36 9.89 2.3 1.730 1.259 2.367 1.266 0.633 Hensei 9 8.67 8.07 3.6 1.008 0.906 3.236 0.010 0.005 Sand Hensei 10 Sand 18.49 16.84 4.7 0.830 1.305 4.113 -1.515 -0.757 Hensei 11 8.29 7.58 4.5 1.142 0.894 3.211 0.156 0.078 Sand Hensei 12 Sand 9.95 9,20 3.9 2.032 0.827 3.565 1.055 0.527 Hensei 13 Sand 11.53 11.36 0.7 1.636 0.902 1.456 2.707 1.353 Hensei 14 15.38 14.72 2.2 0.930 1.315 3.476 -0.310 -0.155 Sand Hensei 15 Sand 15.78 14.84 3.1 1.032 1.394 2.791 0.179 0.089 16 Hensei Sand 15.10 14.50 2.0 0.705 1.364 3.676 -1.515 -0.758 17 Hensei 27.27 25.39 3.6 2.305 1.497 -1.187 8.902 4.451 Sand 18 Alluvium 6.26 5.77 4.1 1.067 0.581 2.366 0.062 0.031 19 Hensei Sand 7.70 - - - 0.505 - - - 21 Hensei 17.46 16.20 3.7 2.806 1.320 0.107 5.340 2.670 Sand 22 Hensei Sand 17.74 17.26 1.4 2.891 1.339 0.116 5.580 2.790 144 Bibliography 145 Bibliography Abbott, P. L, and C. M. Woodruff, Jr., eds, 1986, The Balcones Escarpment-geology, hydrology, ecology and social development in central Texas: Published for the Geological Society of America Annual Meeting, San Antonio, Texas, November, 1986,200 p. Alexander, W. H., Jr., and J. H. Patman, 1969, Groundwater resources of Kimble County, TX, Texas Water Development Board Report 95, 93 p. American Association of Petroleum Geologists, 1973, Geological highway map of Texas, H. B. Renfro Memorial Edition. Amsbury, D. L, R. F. Inden, F. E. Lozo, Jr., C. H. Moore, B. F. Perkins, C. I. Smith, and F. L. Stricklin Jr.,1974, Shallow marine sediments of the early Cretaceous (Trinity) platform of central Texas, a field guide: Society of Economic Paleontologists and Mineralogists, Forty-eighth Annual Meeting, April 4-6,1974, San 1-15. Antonio, TX, p. Anderson, M. P., and W. W. Woessner, 1992, Applied Groundwater Modeling, Simulation of Flow and Advective Transport: Academic Press, 379 p. Ardis, A. F., and R. A. Barker, 1993, Historical saturated thickness of the Edwards-west- Trinity aquifer system and selected contiguous hydraulically connected units, central Texas: U. S. Geological Survey Water Resources Investigations Report 92­ 4125, 2 plates. Ashworth, J. 8., 1983, Ground-water availability of the Lower Cretaceous formations in the Hill Country of south-central Texas: Texas Department of Water Resources Report 273,173 p. facies and flow Back, W. H., 1966, Hydrochemical groundwater patterns in the northern part of the Atlantic coastal plain: U. S. Geological Survey Professional Paper 498-A, 42 p. nature in north- Bain, J. S., 1973, The of the Cretaceous-pre-Cretaceous contact central Texas: Baylor Geological Studies Bulletin 25, 44 p. and E. T. and Barker, R. A., P. W. Bush, Baker, Jr., 1994, Geologic history west-central Texas: hydrogeologic setting of the Edwards-Trinity aquifer system, U. S. Geological Survey Water Resources Investigations Report 94-4039, 51 p. 146 Barker, R. A., and A. F. Ardis, 1992, Configuration of the base of the Edwards-Trinity aquifer system and hydrogeology of the underlying pre-Cretaceous rocks, west-central Texas: U. S. Water Resources Geological Survey Investigations Report 91-4071,25 p. Barnes, V. E., 1981, Geologic Atlas of Texas, Llano sheet: The University of Texas at Austin, Bureau of Economic Geology. Bay, T. A., Jr., 1977, Lower Cretaceous models from Texas and Mexico, in D. G. Bebout, and R. G. Loucks, eds., Cretaceous carbonates of Texas and Mexico, at applications to subsurface exploration: The University of Texas Austin, Bureau of Economic Geology Report of Investigations 89. Bear, J., M. S. Beljin, and R. R. Ross, 1992, Fundamentals of ground-water modeling, EPA/540/S-92/005: Superfund Technology Support Center for Ground Water, Robert S. Kerr Environmental Research Lab, U. S. Environmental Protection 11 Agency, p. Bebout, D. G., and R. G. Loucks, eds, 1977, Cretaceous carbonates of Texas and Mexico, applications to subsurface exploration: The University of Texas at Austin, Bureau of Economic Geology Report of Investigations 89, 332 p. Blum, E., 1982, Soil survey of Kimble County, U. S. Department of Agriculture, Soil Conservation Service, 99 p. Bluntzer, R. L., 1992, Evaluation of the ground-water resources of the Paleozoic and Cretaceous aquifers in the Hill Country of central Texas: Texas Water Development Board Report 339,130 p. Boone, P. A., 1968, Stratigraphy of the basal Trinity Sands (Lower Cretaceous) of central Texas: Baylor Geological Studies Bulletin 15, 64 p. Bush. P. W., A. F. Ardis, and K. H. Wynn, 1993, Historical potentiometric surface of connected the Edwards-Trinity aquifer system and contiguous hydraulically units, Resources west-centra! Texas: U. S. Geological Survey Water Investigations Report 92-4055, 3 sheets. Bush, P. W„ 1986, Planning report for the Edwards-Trinity regional aquifer-system southwest Arkansas: U. S. in central Texas, southeast Oklahoma, and analysis Geological Survey Water Resources Investigation Report 86-4343, 15 p. 147 Campbell, D. H., 1962, Petrography of the Cretaceous Hensell Sandstone, central Texas: M.A. thesis, The University of Texas at Austin, 180 p. Cartwright, L. D. Jr, 1932, Regional structure of the Cretaceous on the Edwards Plateau of southwest Texas: American Association of Petroleum Geologists Bulletin, v. 16, no. 7, p. 691-700. in the crust Chebotarev, 1.1., 1955, Metamorphism of natural waters of weathering: Geochim. Cosmochim. Acta., 8, p. 22-48,137-170, and 198-212. Darcy, H., 1856, Les fontaines publiques deia ville de Dijon: Paris, Victor Dalmont, 647 p. Dixon, W., and B. Chiswell, 1992, The use of hydrochemicai sections to identify recharge areas and saline intrusions in alluvial aquifers, southeast Queensland, 259-274. Australia: Journal of Hydrology, v. 135, p. Dugan, J. T., 1986, Hydrologic characteristics of soils in parts of Arkansas, Colorado, Kansas, Missouri, Nebraska, New Mexico, Oklahoma, South Dakota, and Texas: U. S. Geological Survey Hydrologic Investigations Atlas HA-678, three sheets. W. A., and R. A. Hicks, 1994, Effect of removal of juniperus ashei on Dugas, evapotranspiration and runoff in the Seco Creek watershed: 1994 Annual Report of the Seco Creek Demonstration Project, Texas Agricultural Experiment Station Blackland Research Center, September 1994,12 p. Dunk, W., 1994, rainfall data recorded in the years 1939-1993 by the Dunk family in the Junction area for the National Weather Service. Fetter, C.W., 1988, Applied Hydrogeology, 2nd. ed., Columbus, Ohio, Merrill Publishing Co., 592 p. W. and P. U. Rodda, 1969, Edwards Formation Fisher, L, (Lower dolomitization in a carbonate platform system: American Cretaceous),Texas, Association of Petroleum Geologists Bulletin, v. 53, no. 1, p. 55-72. 148 Fogg, G. E., 1989, Architecture and interconnectedness of geologic media, role of the low-permeability facies in flow and transport, in Hydrogeology of Low Permeability Environments, Special Symposium of the Twenty-Eighth International Geological Congress, July 9-19, 1989: Internationa! Association of Hydrogeologists, Verlag Heinz Heise, p 19-40. Freeze, R.A., and J.A. Cherry, 1979, Ground Water, Englewood Cliffs, New Jersey, Prentice-Hall, 604 p. Golden Software, Inc., 1988, SURFER manual, vers. 4., Golden, Colorado. Hall, D W., and J. L. Turk, 1975, Aquifer evaluation using depositional systems, an example in north-central Texas: Ground Water, v. 13, no. 6, p 472-483. Hammond, W. W., G. Longley, D. J., Pavlicek, and G. B. Ozuna, 1986, Hydrogeology of the Edwards Aquifer: Field Trip Guidebook, Geological Society of America Annual Meeting, San Antonio, TX, 45 p. R. Basic U. S. Heath, C., 1983, ground-water hydrology, Geological Survey Professional Paper 2220, 84 p. Hendricks, L, ed., 1967, Comanchean (Lower Cretaceous) stratigraphy and paleontology of Texas; Permian Basin Section, Society of Economic Paleontologists and Mineralogists Publication 67-8,410 p. Henningsen, R. E., 1962, Water diagenesis in Lower Cretaceous Trinity aquifers of central Texas: Baylor University, Baylor Geological Studies Bulletin 3, 38 p. Water Hill Country Underground Conservation District, 1994, Gillespie County in association with regional water management plan, prepared Blackwell, Lackey and Associates Inc., McGinnis, Lockridge and Kilgore LLP. Hill, R. T., 1901, Geography and geology of the Black and Grand Praries, Texas, with the Cretaceous formations and special reference to detailed descriptions of artesian waters: U. S. Geological Survey, Twenty-First Annual Report, pt. 2, 1899­ 1900,666 p. Hill, R. T., and T. W. Vaughan, 1898, Geology of the Edwards Plateau and Rio Austin and Texas with reference the Grande Plain adjacent to San Antonio, to occurence of underground waters, U. S. Geological Survey Eighteenth Annual Report pt. 11, p.193-321. 149 Holland, P. H., and H. B. Mendieta, 1965, Base-flow studies, Llano River, Texas, quantity and quality: U. S. Geological Survey, March, 1965, 20 p. Hughes, R. J. Jr., 1948, A study of the Lower Cretaceous section near Junction, TX: MA thesis, The University of Texas at Austin, 32 p. Inden, R. F., 1974, Lithofacies and depositional model for Cretaceous a Trinity sequence, central Texas: Louisiana State University, Geoscience and Man 8, p. 37­ 52. Konikow, L. F., and J. Bredehoeft, 1992, Ground-water models cannot be validated: Advances in Water Resources v. 15, p. 75-83. Larkin, T. J., and G. W. Bomar, 1983, Climatic atlas of Texas, Texas Department of Water Resources Publication LP-192,151 p. Lattman, L. H., and R. R. Parizek, 1964, Relationship between fracture traces and v. 73­ the occurance of groundwater in carbonate rocks: Journal of Hydrology, 2, p. 91. rock terranes: LeGrand, H. E., and V. T. Stringfield, 1971, Water levels in carbonate Ground Water, v. 9, no. 3, p 4-10. LeGrand, H. E., and V. T. Stringfield, 1971, Development and distribution of permeability in carbonate aquifers: Water Resources Research, v. 7, no. 5, p. 1284­ 1295. Lloyd, J. W., and J. A. Heathcote, Natural Inorganic Hydrochemistry in Relation to Groundwater, An Introduction: Oxford, Oxford Press, 295 p. Lozo, F. E., and others, 1959, Symposium on the Edwards Limestone in central Texas: The University of Texas at Austin, Bureau of Economic Geology Publication 5905,235 p. Lozo, F. E., and F. L. Stricklin Jr., 1956, Stratigraphic notes on the outcrop basal Cretaceous, central Texas: Transactions, Gulf Coast Association of Geological 67-78. Societies, v. 6, p. relations of northwest Lozo, F. E., 1949, Stratigraphic Fredricksburg limestones, Texas: Shreveport Geological Society Seventeenth Annual Field Trip Guidebook, p 85-91. 150 Maclay, R. W., and L. F. Land, 1989, Simulation of flow in the Edwards aquifer, San Antonio region, Texas, and refinement of and flow U. S. storage concepts: Geological Survey Water Supply Paper 2336, 47 p. Maclay, R. W., and T. A. Small, 1984, Carbonate geology and hydrology of the Edwards aquifer in the San Antonio area, Texas: U. S. Geological Survey Open File Report 83-537, 72 p. Mazor, E., 1991, Applied Chemical and Isotopic Groundwater Hydrology, New York, N. Y., Halsted Press, 274 p. method of converting no-flow cells to variable-head cells for the U.S. Geological Survey modular three-dimensional finite-difference ground-water flow model: U.S. Geological McDonald, M. G., A. W. Harbaugh, B. R. Orr, and D. J. Ackerman,!99l, A Survey Open File Report 91-536, 99 p. M. A. W. A modular three-dimensional finite- McDonald, G., Harbaugh, 1988, Water difference ground-water flow model: U. S. Geological Survey Techniques of Resources Investigations, Book 6, Chapter Al. Menard, J. A., 1995, Bibliography of the Edwards aquifer, Texas, through 1993: U. S. Geological Survey Open File Report 95-336, 75 p. Moore, C. H., Jr., and D. G. Bebout, 1989, Carbonate rock sequences from the Cretaceous of Texas: Field Trip Guidebook T376, American Geophysical Union, Washington D. C. and Moore, C. H., Jr., 1969, ed., Depositional environments depositional history, Lower Cretaceous shallow shelf carbonate sequence, west-centra! Texas: Dallas Annual Meeting of the American Association of Petroleum Geological Society, Geologists and the Society of Economic Paleontologists and Mineralogists, Dallas, Texas, April, 1969, 135 p. Moore, C. H., Jr., 1964, Stratigraphy of the Fredricksburg Division, south-central Economic Geology Report of Texas: The University of Texas at Austin, Bureau of Investigations 52, 48 p. Murray, C. R., and E. B. Reeves, 1972, Estimated use of water in the United States in 1970: U. S. Geological Survey Circular 676, 37 p. 151 Myers, B. N., 1969, Compilation of results of aquifer tests in Texas, Texas Water Development Board Report 98, 532 p. Payne, J. Hopkins, 1982, Sedimentation and pedogenesis of the Lower Cretaceous Hensell Formation, central Texas: M.A. thesis, The University of Texas Austin, at 136 p. Pearson, F. J., and P. L. Rettman, 1976, Geochemical and isotopic analysis of waters associated with the Edwards limestone aquifer, central Texas: U. S. Geological Survey and the Edwards Underground Water District, 34 p. Piper, A. M., 1953, A graphic procedure in the geochemical interpretation of water 914-923. analysis: American Geophysical Union Transactions, v. 25, p. Rapp, K. 8., 1988, Groundwater recharge in the Trinity aquifer, central Texas: Baylor Geological Studies Bulletin 46, Baylor University, 34 p. Reddi, L. N., 1990, Potential pitfalls in using groundwater models, transferring models to users: American Water Resources Association, Nov. 1990, p. 131-140. Rose, P. R., 1972, Edwards Group, surface and subsurface, central Texas: The University of Texas at Austin, Bureau of Economic Geology Report of Investigations 74,198 p. Sami, K., 1992, Recharge mechanisms and geochemical processes in a semi-arid 27­ sedimentary basin, Eastern Cape, South Africa: Journal of Hydrology, v. 139, p. 48. Sharp, J. M. Jr., 1990, Stratigraphic, geomorphic, and structural controls on the Edwards Aquifer, Texas, USA, in E. S. Simpson and J. M. Sharp Jr., eds., Selected International Association papers on hydrogeology: of Hydrogeologists, Heise, Hannover, v. 1, p. 67-82. Stiff, H. A., Jr., 1951, The interpretation of chemical water analysis by means of patterns: Journal of Petroleum Technology 3, no. 10, sec. 1, and no. 15-16, sec. 2, 3. L. C. I. and F. E. Lozo, 1971, of Lower Stricklin, F. Jr., Smith, Stratigraphy Cretaceous Trinity deposits of central Texas: The University of Texas at Austin, Bureau of Economic Geology Report of Investigations 71,63 p. 152 Texas Natural Resources Information Service (TNRIS), 1994, Streamflow data for U. S. Geological Survey stream gauge # 08150000, Llano River at Junction. Toth, J., 1963, A theoretical analysis of groundwater flow in small drainage basins, Journal of Geophysical Research v. 67, p. 4795 4812. - Toth, J., 1962, A theory of groundwater motion in small drainage basins in Central Alberta, Canada: Journal of Geophysical Research v. 67, p. 4375-4387. U. S. Geological Survey, 1989, National Water Summary 1988-89, Floods and 513-520. Droughts, Texas, U. S. Geological Survey Water Supply Paper 2375, p. Walker, L., 1979, Occurence, availability and chemical quality of ground water in the Edwards Plateau region of Texas: Texas Department of Water Resources Report 235, 337 p. Walton, W. C., 1962, Selected analytical methods for well and aquifer evaluation, State of Illinois Water Survey Bulletin 49, 81 p. Wang, H. F., and M. P. Anderson, 1982, Introduction to Groundwater Modeling, Finite Difference and Finite Element Methods, San Francisco, CA, W. H. Freeman and Co., 237 p. Wermund, E. G., J. C. Cepeda, and P. E. Luttrell, 1978, Regional distribution of fractures in the southern Edwards Plateau and their relationship to tectonics and caves: The University of Texas Austin, Bureau of Economic Geology Geological at Circular 78-2,15 p. Wood, W.W., 1981, Guidelines for collection and field analysis of ground-water samples for selected unstable constituents, U.S. Geological Survey Techniques of Water Resources Investigations, Chapter d2, 24 p. Land in the Lake Travis Woodruff, C. M., 1975, capability vicinity, Texas, a practical guide for the use of geologic and engineering data: The University of Texas at Austin, Bureau of Economic Geology Report of Investigations 84, 37 p. Woodruff, C. M., and R. M. Slade Jr., 1985, Hydrogeology of the Edwards Aquifer, Barton Springs segment, Travis and Hays Counties, Texas: Austin Geological Society Guidebook 6, 96 p. 153 Yelderman, J. C. Jr., coordinator, 1987, Hydrogeology of the Edwards Aquifer, northern Balcones and Washita Prarie Austin Geological Society segments: Guidebook 11,91 p. 154 The vita has been removed from the digitized version of this document.