Loading...
HomeMy WebLinkAboutNC0038377_App B GW Modeling Rpts_20160229GROUNDWATER FLOW AND TRANSPORT MODELING REPORT FOR MAYO STEAM ELECTRIC PLANT, November 8, 2015 Prepared for SynTerra 148 River Street Greenville, SC 29601 Investigators Lawrence C. Murdoch, Ph.D. Scott E. Brame, M.S. Ronald W. Falta, Ph.D. Regina Graziano, M.S. Contents 1.0 Introduction............................................................................................................. 5 1.1 General Setting and Background........................................................................... 5 1.2 Study Objectives................................................................................................... 6 2.0 Conceptual Model................................................................................................... 7 2.1 Aquifer System Framework.................................................................................. 7 2.2 Groundwater Flow System.................................................................................... 8 2.3 Hydrologic Boundaries...................................................................................... 9 2.4 Hydraulic Boundaries............................................................................................ 9 2.5 Sources and Sinks................................................................................................ 10 2.6 Water Budget....................................................................................................... 10 2.7 Modeled Constituents of Interest........................................................................ 10 2.8 Constituent Transport .......................................................................................... 11 3.0 Computer Model................................................................................................... 12 3.1 Model Selection................................................................................................... 12 3.2 Model Description............................................................................................... 12 4.0 Groundwater Flow and Transport Model Construction .................................. 13 4.1 Model Domain and Grid..................................................................................... 13 4.2 Hydraulic Parameters.......................................................................................... 15 4.3 Flow Model Boundary Conditions...................................................................... 15 4.4 Flow Model Sources and Sinks........................................................................... 16 4.5 Flow Model Calibration Targets......................................................................... 18 4.6 Transport Model Parameters............................................................................... 18 4.7 Transport Model Boundary Conditions............................................................... 19 4.8 Transport Model Sources and Sinks................................................................... 20 4.9 Transport Model Calibration Targets.................................................................. 20 5.0 Model Calibration to Current Conditions.......................................................... 21 5.1 Flow Model Residual Analysis........................................................................... 21 5.2 Flow Model Sensitivity Analysis........................................................................ 24 5.3 Transport Model Calibration and Sensitivity...................................................... 24 5.4 Result of the Transport Simulation.................................................................. 25 6.0 Predictive Simulations of Corrective Action Scenarios ..................................... 27 6.1 CAP 1 -No Action............................................................................................... 27 6.2 CAP2 - Capping Ash in Place............................................................................. 28 6.3 CAP3 - Complete Ash Removal......................................................................... 30 7.0 References.............................................................................................................. 31 Tables................................................................................................................................ 32 Figures............................................................................................................................... 41 LIST OF TABLES Table 1. Observed and computed heads for the calibrated flow model. Table 2. Calibrated hydraulic parameters. Table 3. Recharge in different zones in the model. Table 4. Flow parameter sensitivity analysis Table 5. Comparison of observed and simulated boron concentrations (ug/L) in monitoring wells. Table 6. Comparison of observed and simulated arsenic concentrations (ug/L) in monitoring wells. LIST OF FIGURES Figure 1. Map of the Mayo Steam Electric Plant showing ash basin, hydrologic features and boundary of model region. Figure 2. Air photo of the Mayo Steam Electric Plant showing ash basins and hydrologic features and boundary of model region (yellow). Figure 3. Fence diagram of the 3D hydrostratigraphic model (Solids) used to generate hydrostratigraphy. Figure 4a. Perspective of site looking SW along Crutchfield Branch showing computational grid in cross section. Figure 5. Distribution of recharge flux used in the Mayo model. Scale in ft/d. Figure 6. Surface water features and wells in the vicinity of the Mayo Plant. Figure 7. Hydraulic conductivity (ft/d) distributions in grid layers in the Mayo model. Figure 8. Zones of different hydraulic conductivity enlarged and condensed from Figure 7. Figure 9. Hydraulic head computed at steady state as a function of the observed head in monitoring wells in the vicinity of the Mayo Plant. Figure 10. Steady state hydraulic head distribution predicted in top of bedrock (layer 8 of the model) in the Mayo model. Figure 11. Steady state hydraulic head distribution in predicted in top of bedrock (layer 8 of the model) in the Mayo model. Figure 12. Steady state hydraulic head distribution in the vicinity of the ash basin (layer 4 of the Mayo model). Figure 13. Concentrations of boron in ash used to specify concentrations in the Mayo model. Figure 14. Concentration of boron predicted to occur in Oct 2015 in different layers of the Mayo model. Figure 15. Concentrations of boron predicted to occur in Oct 2015 in different layers. Figure 16. Map and cross -sections describing design of the Cap -in -Place design used to develop the model of this scenario. Figure 17. Distribution of recharge in simulations of corrective actions. Figure 17a. Hydraulic head calculated in the vicinity of the ash basin for the different corrective actions. Figure 18-73. Distributions of concentrations of boron and arsenic at 2015, 2020, 2030, 2045, layers in the model, and for different corrective actions. 1.0 INTRODUCTION Duke Energy Progress, LLC. (Duke Energy) owns and operates the Mayo Steam Electric Plant (Mayo Plant) located near Roxboro, in Person County, North Carolina. The Mayo Plant began operations in the 1983. In 2013, the Mayo Plant converted from a wet to dry ash system. Consequently, 90 percent of currently generated CCR is dry. Mayo Plant is a single unit facility. Coal combustion residuals (CCRs) have been stored in one ash basin located northwest of the Plant. The ash basin is approximately 144 acres in size, is constructed with an earthen dam, and contains approximately 6,900,000 tons of CCR (Duke Energy, October 31, 2014). 1.1 General Setting and Background The Mayo Plant ash basin is located in the northern portion of the Plant property, northwest of the railroad line that effectively bisects the plant. The ash basin is impounded by an earthen embankment system approximately 2,300 feet long, with a height of 110 feet. The entire basin area is approximately 144 acres and contains approximately 6,900,000 tons of CCR material (Duke Energy, October 31, 2014). The ash basin is the only ash storage area at the site with the exception of the recently operational CCR monofill. The ash basin is approximately half open water and half ash. Heavily wooded land surrounds the ash basin with the exception of the wastewater treatment facility on the southeast side of the ash basin. The Mayo Plant ash basin was constructed with two engineered toe drains located at the base of the dam. In addition, ash basin water is diverted into a holding lagoon or forebay for water quality treatment. Water flows from the ash basin into the forebay through a 48-inch pipe, riser, and decant pipe through the forebay embankment. Discharge from the forebay is controlled by a concrete overflow weir. Discharge from the forebay passes over the weir and flows through NPDES Outfall 002, through a discharge canal and into Mayo Lake. The Mayo Plant is in the eastern Piedmont Region, which is underlain by metamorphic and igneous rocks (Trapp and Horn, 1997). The Piedmont region of North Carolina is characterized by gently rolling hills and ridges, however the topography around the Mayo Plant is steeper with incised streams and deeply cut ridges. The ash basin was created by impounding Crutchfield Branch, a small stream that flows to the north. Mayo Creek, the drainage to the east, was impounded to create Mayo Lake. In general, the topography slopes away from the entrance road and the power plant area towards Mayo Lake. The power plant area is situated at an approximate elevation of 520 feet MSL and Mayo Lake is at an elevation of about 435 feet MSL, a vertical difference (relief) of 85 feet. The area is underlain by metamorphic rock, which is fractured and weathered at shallow depths. Saprolite is up to 25 feet thick, but is absent locally. It is unsaturated in the uplands, but is saturated at lower elevations streams. A transition zone of partially weathered rock underlies the saprolite and is generally continuous throughout the Mayo Plant area. The transition zone at the Mayo Plant is partially weathered rock that is gradational between saprolite and competent bedrock. It is as thick as 22 ft in the vicinity of the Mayo Plant. The degree of weathering decreases with depth (SynTerra, 2015). The water table in the upland areas is typically in the fractured rock. Alluvium occurs along Crutchfield Branch and Mayo Creek. 1.2 Study Objectives The purpose of this study is to predict the groundwater flow and constituent transport that will occur as a result of different possible corrective actions at the site. The study consists of three main activities: develop a calibrated steady-state flow model of current conditions, develop a historical transient model of constituent transport that is calibrated to current conditions, and create predictive simulations of the different corrective action options. 2.0 CONCEPTUAL MODEL The site conceptual model for the Mayo Plant is primarily based on the Comprehensive Site Assessment Report (CSA Report) for the Mayo Site (SynTerra, 2015). The CSA report contains extensive detail and data related to most aspects of the site conceptual model. 2.1 Aquifer System Framework The aquifer system at the site is unconfined and includes three main hydrostratigraphic units, a saprolite/transition zone, upper bedrock, and lower bedrock (Legrand, 1988). The saprolite/transition zone consists of partially to thoroughly weathered (SynTerra, 2015). It is underlain by fractured metamorphic rock. The degree of fracturing is spatially variable and generally decreases downward. Vertical and horizontal fractures zones can cause localized zones of high permeability within the rock (Legrand, 1988; Miller, 1990). The permeability is moderate in many of the bedrock wells, and it is inferred that the fracture density and hydraulic conductivity decrease downward (Legrand, 1988). The saprolite/transition zone is saturated in the vicinity of streams and lakes where groundwater is discharging, but it is unsaturated in most upland areas. The water table occurs in the fractured bedrock in most upland areas. Saprolite thickness ranges up to 66 feet, but it is less than 25 ft thick in most locations. Alluvium was observed at two locations along the Crutchfield Branch with a maximum thickness of 7.5 feet. Ash is saturated in the ash basins. The water level in the ash basin is at approximately 480 ft and the ash surface rises up to approximately 490 ft at the southern end of the basin. The elevation of Crutchfield Branch stream is approximately 390 ft at the ash basin dam. Hydraulic conductivity was determined in the field using slug tests. The hydraulic conductivity of the saprolite was 0.02 to 0.45 ft/day with a geometric mean of 0.2 ft/d. The transition zone ranged from 0.02 to 3 ft/d with a geometric mean of 0.015 ft/d. These measurements reflect the variability of the transition zone, where hydrologic properties influenced by the localization of clays and fragments of unweathered rock. The hydraulic conductivity of the bedrock spanned a broad range, from 0.003 ft/d to 5.6 ft/d, according to the results from 29 slug tests on 15 wells. The high values were probably measured at wells that intersect fractures zones in the rock. The hydraulic conductivity of the ash was measured by conducting a total of 5 slug tests at 4 wells completed in the ash. Hydraulic conductivity spanned 2 orders of magnitude, from 0.05 to 4 ft/d, and the geometric mean was 0.4 ft/d. 2.2 Groundwater Flow System The groundwater system is recharged by rain water and from water that infiltrates through the ash basin. The average value of recharge was estimated from the map of recharge in North Carolina by Haven (2003) and from analyzing stream hydrographs. A shapefile of the recharge map by Haven (2003) was enlarged and features of the site were superimposed on it. Colors on the map were compared to colors on the legend because quantitative data were unavailable from the file. This indicated that recharge was in the range of 4-12 inch/yr (10-30 cm/yr) in the watershed draining into the ash basin (Figure 1). The flow in Hyco Creek was obtained from measurements made at the gauging station USGS 02077200 near Leasburg, NC. The gauging station is in Caswell County approximately 20 mi SW of the site and it measures flow from a watershed approximately 45.9 mil to the south. Hyco Creek flows into Hyco Lake. The analysis was conducted on I I years of data starting in January, 2002. The hydrograph was analyzed by separating stormflow and baseflow from the hydrograph using the method described by the Institute of Hydrology (1980). This method of hydrograph separation is widely used by the USGS and others. The separated hydrograph was analyzed using methods described by Mau and Winter (1997), and Rutledge and Mesko (1996) to estimate the recharge required to produce the observed baseflow. Recharge was estimated on a monthly basis and then averaged over the time period of the dataset. This resulted in an estimate of recharge that ranges from 3 to 7 inch/yr, depending on how the recharge is assumed to occur between baseflow turning points. Recharge estimated using the hydrograph from Hyco Creek was generally less than that shown on from the map by Haven (2003), although the ranges from the two methods overlap. Both of these methods of estimating recharge have advantages and disadvantages, so it was assumed that the recharge the upland areas was 0.0018 ft/d (approximately 6.5 in/yr). Further, it was assumed that recharge was negligible in the vicinity of the plant, the lined treatment lagoon, surface water, and the low permeability dams. Specific values assumed for the recharge are given later. The Mayo Plant ash basin occupies the former stream valley of Crutchfield Branch, a headwater stream flowing northeast. The basin was created by building a dam across the stream valley. A small stream flows into the ash basin, and groundwater is inferred to discharge into the basin from uplands to the west, south and east of the basin. The surface watershed that drains into the basin is bounded on the east and south by a divide with drainage into Mayo Lake. It is bounded on the west by a divide along US Highway 501 (SynTerra, 2015). The ash basin is drained by Crutchfield Branch stream, which flows north off the Mayo Plant property into Virginia (Figures 1 and 2). Water levels in the wells completed in the ash basin were similar to the level of surface water behind the dam. The water levels ranged from 480 ft to 485 ft in the ash basin wells, with heads that increased toward the south where the headwaters of Crutchfield Branch enter the basin (Figure 1). The distribution of hydraulic head suggests that ground water is likely flowing into the upstream end of the ash basin from the uplands to the west, south and east. The water level near the dam is several 1 Os of ft higher than the water level in wells along Crutchfield Branch. It is inferred that this large head difference drives groundwater flow north from the ash basin. Approximately 21 water supply wells have been identified within one half mile of the Mayo Plant (SynTerra, 2014). Most of the private wells are either northwest or south of the Mayo Plant. The wells were sampled by NCDNER for chemical analysis, but measurement data on the discharge rate from the wells was unavailable. The average daily water use is 80 to 100 gals per person, so a well providing water for a family of four people would be pumped at approximately 350 gals per day. 2.3 Hydrologic Boundaries The major discharging locations for the shallow water system serve as hydrologic boundaries to the shallow groundwater system. These include lakes and streams. 2.4 Hydraulic Boundaries The shallow groundwater system does not appear to contain impermeable barriers or boundaries in the study area. The degree of fracturing, and thus the hydraulic conductivity, is expected to decrease with depth in metamorphic rock. This will result in blocks of unfractured rock where the hydraulic conductivity is quite low. However, isolated fractures may occur that result in large local hydraulic conductivities, and the locations of these fractures is difficult to predict. It was assumed that the rock was impermeable below mean sea level, and this elevation was used as the bottom of the model. 2.5 Sources and Sinks Recharge is the major source of water in the uplands and ash basins. Most of the groundwater discharges to streams and lakes, as outlined above. Groundwater flows through the ash basin, so it acts as both a source of, and sink for groundwater. Domestic water supply wells act as sinks to groundwater to the south and north of the plant. (SynTerra, 2015). Screen elevations and pumping rates from most of these wells are unknown. 2.6 Water Budget Over the long term, the rate of water inflow to the study area is equal to the rate of water outflow from the study area. Water enters the groundwater system through recharge, and leaves through discharge to surface water and wells. 2.7 Modeled Constituents of Interest Antimony, arsenic, barium, cobalt, iron, manganese, pH, thallium, total dissolved solids (TDS), and vanadium were detected in the ash basin (pore) water, near the bottom of the ash, at concentrations greater than 2L or IMAC. Of these constituents, all but arsenic, barium, and TDS were also detected in Site background wells upgradient of the ash basin at concentrations greater than 2L or IMAC. Neither arsenic nor barium was detected greater than 2L or IMAC outside of the ash basin. Cobalt, iron, manganese, and vanadium are commonly detected in shallow groundwater in the Piedmont of North Carolina. Site background concentration ranges for some of these constituents are available from routine monitoring of the upgradient compliance boundary monitoring wells and newly installed background wells. Boron, cobalt, and TDS are the only COIs in groundwater detected downgradient of the ash basin (SynTerra, 2015). Cobalt is common in background wells, so the occurrence downgradient is difficult to distinguish from background. TDS is also difficult to distinguish from background. Statistical analyses of the occurrence of cobalt and TDS would help identify whether these compounds could be distinguished from background. The COIs selected for modeling at the Mayo site are boron and arsenic. Boron is the primary COI, but arsenic was also included because it is present in the ash at concentrations several times greater than the 2L standard and it has a relatively low 2L standard. Other constituents were not used in the modeling exercise for one or more of the following reasons: 1) concentrations in the ash pore water do not greatly exceed likely background levels; and 2) there is no discernable plume of the constituent extending downgradient from the ash basin. 2.8 Constituent Transport The COIs that are present in the coal ash dissolve into the ash pore water. As water infiltrates through the basins, water containing COIs can enter the groundwater system through the bottom of the ash basins. Once in the groundwater system, the COIs are transported by advection and dispersion, subject to retardation due to adsorption to solids. If the COIs reach a hydrologic boundary or water sink, they are removed from the groundwater system, and they enter the surface water system, where in general, they are greatly diluted. At this site, boron is the primary constituent that is migrating from the ash basin. 3.0 COMPUTER MODEL 3.1 Model Selection The numerical groundwater flow model was developed using MODFLOW (McDonald and Harbaugh, 1988), a three-dimensional (3D) finite difference groundwater model created by the United States Geological Survey (USGS). The chemical transport model is the Modular 3-D Transport Multi -Species (MT3DMS) model (Zheng and Wang, 1999). MODFLOW and MT3DMS are widely used in industry and government, and are considered to be industry standards. The models were assembled using the Aquaveo GMS 10.0 graphical user interface (http://www.aquaveo.com/). 3.2 Model Description MODFLOW uses Darcy's law and the conservation of mass to derive water balance equations for each finite difference cell. MODFLOW considers 3D transient groundwater flow in confined and unconfined heterogeneous systems, and it can include dynamic interaction with pumping wells, recharge, evapotranspiration, rivers, streams, springs, lakes, and swamps. Several versions of MODFLOW have been developed over the years. This study uses the MODFLOW-NWT version (Niswonger, et al., 2011). The NWT version of MODFLOW provides improved numerical stability and accuracy for modeling problems with variable water tables. That improved capability is helpful in the present work where the position of the water table in the ash basin can fluctuate depending on the conditions under which the basin is operated and on the corrective action activities. MT3DMS uses the groundwater flow field from MODFLOW to simulate 3D advection and dispersion of the dissolved COIs including the effects of retardation due to COI adsorption to the soil matrix. 4.0 GROUNDWATER FLOW AND TRANSPORT MODEL CONSTRUCTION The flow and transport model for this site was built through a series of steps. The first step was to build a 3D model of the site hydrostratigraphy based on field data. The next step was determination of the model domain and construction of the numerical grid. The numerical grid was then populated with flow parameters, which were adjusted during the steady-state flow model calibration process. Once the flow model was calibrated, the flow parameters were used to develop a transient model of the historical flow patterns at the site. The historical flow model was then used to provide the flow field for the transient constituent transport simulations. Calibration of the transport model required some adjustments in order to reproduce the observed boron plume. This resulted in a second iteration of flow model calibration, so that the calibrated flow model matches the observed heads, and the transient flow and transport model reproduce the observed boron plume. 4.1 Model Domain and Grid The first steps in the model grid generation process were the determination of the model domain, and the construction of a 3D hydrostratigraphic model. The model has dimensions of approximately 3 miles by 3 miles. The model domain was rotated 31 ° clockwise so boundaries of the model were parallel with the ash basin dam. The shortest distance between the ash basin and a model boundary was approximately 1 mile. The ground surface of the model was interpolated from USGS NED n37w079 1/3 arc -sec 2013 1 degree IMG dataset obtained from http://viewer.nationalmgp.gov/viewer/. The elevations for the top of the ash basin were modified using more recent surveying data. The hydrostratigraphic model consists of six units: Ash, Saprolite, Transition Zone, Upper Bedrock, Middle Bedrock, and Lower Bedrock. The hydrostratigraphic model was developed using "Solids" in GMS (Figure 3). Five solids were created and then subdivided after the computational mesh was developed. The solids include Ash, Saprolite, Transition Zone, Fractured Rock, and Rock. The lower contact between the ash basin and the underlying saprolite was assumed to be the ground surface prior to construction of the ash basins. An electronic file describing this surface was created by digitizing a preconstruction topographic map. The digitized points were interpolated to create a continuous surface representing the preconstruction ground surface, and this was used as the contact between the ash and the underlying saprolite. This created a surface that was consistent with the borehole observations except at A13MW-4 where the contact was 20 ft lower than expected. This may have resulted from construction -related excavation that was not recorded on the pre -construction topo map. The bottom contact between the ash and underlying weathered rock was lowered in the vicinity of ABMW-4 as a result. The lateral extent of the ash was determined from aerial photographs and from maps in the CSA report (SynTerra, 2015). The contacts between the saprolite, transition and underlying bedrock were determined by interpolating data measured in borings described in the CSA report and historical data. This produced two isopach maps, one showing the thickness of the saprolite and the other showing the thickness of the partially weathered rock in the transition zone. The interpolated isopach surface for the saprolite was subtracted from the ground surface to create a surface marking the contact between the saprolite and transition zone. The isopach map for the transition zone was subtracted from the surface describing the saprolite-transition zone contact. This created a surface that was used as the contact between the transition zone and the underlying fractured rock. The methodology outlined above for creating a geologic model was done so the interpolated contacts would follow the ground surface between boreholes, which is consistent with the expectations based on the hydrogeology of the Piedmont region (e.g. LeGrand, 1988; Miller, 1990). A contact between the fractured and relatively unfractured bedrock was assumed to occur 100 ft below the bottom of the transition zone. The numerical model grid consists of 15 layers representing the hydrostratigraphic units. The model grid was set up to conform to the contacts from the solids. The model grid layers correspond to the solids as follows: H drostrati ra hic layer Grid layer Ash 1-4 Saprolite 5 Transition zone 6-7 Upper fractured rock 8-10 Middle fractured rock 11-12 Rock 13-15 Grid layers 1-4 were set as inactive outside of the region of the ash basin as determined from aerial photos and the CSA report. Grid layers 1-15 were set as inactive in the eastern, southern, and western corners of the model domain (Figure 1). The numerical grid consists of rectangular blocks arranged in columns, rows and layers. There are 171 columns, 232 rows, and 15 layers (Figure 4). The maximum width of the columns and rows is 100 ft. The size of the grid blocks is approximately 50 ft in the vicinity of the ash basins. The horizontal dimension of some of the grid blocks is as small as 25 ft in the vicinity of the dams. 4.2 Hydraulic Parameters The horizontal hydraulic conductivity and the horizontal to vertical hydraulic conductivity anisotropy ratio (anisotropy) are the main hydraulic parameters in the model. The distribution of these parameters is based primarily on the model hydrostratigraphy, with some local variations. Many of the hydraulic parameter distributions in the model were uniform throughout a model layer. Initial estimates of parameters were based on literature values, results of slug and core tests, and simulations performed using a preliminary flow model. The hydraulic parameter values were adjusted during the flow model calibration process described in Section 5.0 to provide a best fit to observed water levels in observation wells. 4.3 Flow Model Boundary Conditions The outer lateral boundary conditions for the saprolite is almost entirely constant head, with small areas of no -flow locally. Boundaries on the east side the model include parts of Mayo Lake, which were held at specified head. The boundaries on the south and east and north sides of the model are independent of definitive hydrologic features. A specified head boundary condition with the head set in the middle of the transition zone was used along these boundaries. This boundary condition forces the water table to be in the transition zone along these boundaries, which is a reasonable approximation of the expected conditions. The specified head boundary condition extends along the upland areas, but it is terminated within a few hundred ft the locations of streams or lakes. This is because streams or lakes that intersect the external boundary are defined by their own boundaries conditions (as either constant head or drain -type boundaries). This creates short intervals of no -flow conditions between streams or lakes and the uplands. The constant head boundary condition along the outside of the model was assigned to Layer 7 in the grid, which is the lowest layer in the transition zone. No -flow conditions were assumed along the boundary beneath Layer 7. 4.4 Flow Model Sources and Sinks The flow model sources and sinks on the interior of the model consist of recharge, lakes, streams, and groundwater pumping. Recharge is a key hydrologic parameter in the model (Figure 5). As described in Section 2.2, the recharge rate for upland areas in the vicinity of the Mayo Plant was assumed to be 0.0018 ft/d (6.5 inches/year). The recharge on exposed ash was assumed to be 0.0018 ft/d, the same as in upland areas. This is because the shallow water table would have increased evaporation, while the lack of vegetation would have decreased ET on the ash basin compared to the upland area. As a result, without field data it was difficult to assess how the recharge on the ash basin would have differed from the recharge on the uplands. The recharge was assumed to be zero at the ash basin lake and on Mayo Lake. The recharge rate on the Mayo Plant was set to 0.0001 ft/d, due to the large areas of roof and pavement. The recharge beneath a lined treatment pond north of the plant site was assumed to be 0.0001 ft/d (Figure 5). Recharge on the dams themselves was set to 0.0001 ft/d, which is consistent with the low hydraulic conductivity assumed for the dams. Recharge was not adjusted much during the model calibration process, but it is included in the sensitivity analysis. The reason for not including recharge as a calibration parameter is that for steady-state flow, the hydraulic heads are determined primarily by the ratio of recharge to hydraulic conductivity, so the two parameters are not independent. In situations where the groundwater discharges to a flow measuring point (for example a gauged stream in a watershed), the flow measurement can be used to calibrate the recharge value allowing both the recharge rate and the hydraulic conductivity to be simultaneously calibrated. However, no streams were gauged at the Mayo site, so this was impossible for this model and the recharge was fixed. Lakes were represented as specified head boundaries with the head set to their stage (Figure 6). This includes Mayo Lake and the ash basin lake. The stage of the ash basin was set to 480 ft based on Lidar data and a surveying point. The stage of Mayo Lake was 432 ft. Streams were represented as type 3 boundary conditions, called "drains" in MODFLOW (Figure 6). The elevation of the streams was set to the ground surface elevation determined from the Lidar data. The drain conductance was set to 100 ft2/day, a relatively large value that will cause negligible head loss, and was not adjusted during calibration (Figure 6). The ash basins were represented by simulating the observed surface water as specified head and applying recharge on the exposed ash. This approach treats the ash basins in the same way as other hydrogeologic components in the model, and it was selected as the best approach to characterize current conditions. The hydrologic conditions when the ash was below the level of the water early in the life of the basin differed from the current conditions. However, the hydraulic heads in the ash today are only slightly above the level of the water. It is likely that the ground water flow system created by the ash basin filled only with water would have been essentially the same as the basin containing ash as it does today. As a result, the hydrologic conditions in the ash basin will be assumed to be constant through the life of the basin. The outflow channel was represented as specified head, with the stage set to elevations measured by surveying. This was done because this engineered channel could exchange water with the groundwater system, and the head was measured in the field. Relatively little information was available about the wells in the model area. The location of water supply wells in the model area are shown as yellow squares in Figure 6 (SynTerra, 2015). Most of the wells are probably open holes in the upper few 100 ft of bedrock. However, it is common for drillers in the Piedmont to extend wells to depths of several 100 ft in an effort to intersect permeable fractures and create more productive wells. As a result, the depth of the wells probably ranges from 150 ft to 600 ft. The wells are assumed to be screened in grid layers 10-12 in the model. The pumping rates from the wells were unknown, but it was assumed that the wells were pumped at about 350 gallons/day, which is an average water use for a family of four. 4.5 Flow Model Calibration Targets The steady state flow model calibration targets were 47 water level measurements made in observations wells in June, 2015. The flow model calibration target wells are listed in Table 1. In general, wells with a D designation at the end of the name are screened in the transition zone, those with a BR designation are screened in the upper bedrock, and those with an S designation are in saprolite. Wells with ABMW are screened in ash. 4.6 Transport Model Parameters The transport model uses a MODFLOW simulation to provide a groundwater velocity field. The transport simulation was started April, 1983, and it continued through Sept, 2015. The Mayo Plant began operations in 1983, and it was assumed that the basin was filled with water at this time. The flow model assumes that the ash basin filled with water quickly and the heads were maintained at the same level as they are today. As a result, a steady state flow field calculated by calibration to the current conditions was used to simulate water flow during transport. The key transport model parameters (besides the flow field) are the constituent source concentration in the ash basin, and the constituent soil -water distribution coefficients (Kd). Secondary parameters are the longitudinal, transverse, and vertical dispersivity, and the effective porosity. The constituent source concentrations were estimated from recently measured ash pore water concentrations in monitoring wells (SynTerra, 2015). Linear adsorption Kd values for Mayo Plant COIs were measured in the laboratory using core materials from the coal ash and native aquifer materials (Langley, et al., 2015). In general, the measured Kd values for the constituents were highly variable, and the variability within a given material type was larger than the variability between different materials. In light of the variability of the measured Kd values, it was decided to use Kd as a calibration parameter. The initial value used in calibration was 5 mL/g, which is the low end of the range measured by Langley et al. (2015). It was found during the transport model calibration that Kd = 0.12 mL/g for boron. The Kd value used for arsenic in the simulations was 50 mL/gm. This is on the low end of the range of Kd values determined in the laboratory by Langley et al. (2015), which ranged from 45 to 1,800 mL/g in batch, and from 6 to 3,349 mL/g in column experiments. Arsenic was only detected in significant concentrations below the ash basins where the uncertainty in the distance between the bottom of the ash and the screen on the monitoring well was relatively large. As a result of this uncertainty and the limited number of data points, it was decided to use a laboratory value for Kd. A value that was on the low end of the range that included both column and batch data was selected so that the transport calculations for arsenic would predict the most rapid transport feasible within the range of the data. This was done to bound the maximum transport distances for arsenic. The longitudinal dispersivity was assigned a value of 20 ft, the transverse dispersivity was set to 2 ft, and the vertical dispersivity was set to 0.2 ft. The soil dry bulk density was set to 1.6 g/mL. The effective porosity was assumed to decrease with depth from 0.3 in the ash, to 0.2 in the saprolite and transition zone to 0.001 in the deep rock. The effective porosity was assumed to decrease with depth based on the hydrogeologic conceptual model. It was assumed the effective porosity was uniform within a grid layer and was distributed according to the following table: Layer Effective porosity 1-4 0.3 5-7 0.2 8 0.05 9-10 0.01 11-15 0.001 4.7 Transport Model Boundary Conditions The transport model boundary conditions are no flow on the exterior edges of the model. The infiltrating rainwater is assumed to be clean, and enters with zero concentration from the top of the model. Contaminants are assumed to leave the model when they arrive at a drain, or are removed by flow that enters a constant head boundary. The initial condition for the current conditions transport model is one of zero concentration of COIs in groundwater. No background concentrations are considered. The concentration in the ash basin is assumed to rapidly increase to the observed concentrations at the start of the simulation. 4.8 Transport Model Sources and Sinks The ash basins are the source of COIs in the model. These sources are simulated by holding the COI concentration constant in cells located inside the ash basins (Figure 13). This allows infiltrating water to carry dissolved constituents from the ash into the groundwater system. Chemical analyses from four wells were used to characterize the distribution of COI concentration within the ash basins. The concentration observed in the wells was assumed to represent the concentration in the vicinity of the well throughout the simulation. This resulted in a patch -like distribution of concentration within the ash basins (Figure 12). Layer 5 in the vicinity of ABMW-4 was assumed to be ash, as described above. Layer 5 is saprolite elsewhere in the model, but the contact between saprolite and ash was approximately 20 ft lower than estimated from the pre -ash -basin topography. It seems likely that this discrepancy occurred because of excavation during construction. The concentration in the vicinity of ABMW-4 in Layer 5 was held constant to resemble the behavior of ash. This created a zone of relatively high concentration in Layer 5. The outflow channel flowing south from the ash basin is represented as constant head and it is possible that water could flow from this engineered feature to the ground water. It was assumed that the concentration in the surface water at this location was zero. This assumption had no bearing on the results, however, because the outflow channel gained ground water. The transport model sinks are potentially lakes and streams. As groundwater enters these features, it is removed along with any dissolved constituent mass. Similarly, if water containing a constituent were to encounter an extraction well, the constituent would be removed with the water. 4.9 Transport Model Calibration Targets The transport model calibration targets are COI concentrations measured in monitoring wells in June, 2015 (SynTerra, 2015). 5.0 MODEL CALIBRATION TO CURRENT CONDITIONS 5.1 Flow Model Residual Analysis The flow model was calibrated in stages starting with a model that assumed homogeneous conditions in most formations. In general, calibration was done by seeking the simplest configuration of parameters that matched the observed hydrogeologic conditions and the assumed or observed geologic conditions. Many of model layer properties were homogeneous. Several heterogeneities were assumed to improve the fit between the simulated and observed heads and concentrations. The calibration was initiated using the geologic model to define the geometry of hydrogeologic units and assigning hydraulic conductivities typical of the region. PEST was then use to minimize the residual between predicted and observed heads. This resulted in reasonably close matches, but there were several wells where the simulation significantly over- or under - predicted the heads. The next step was to infer heterogeneities that could reduce the residuals. The model over -predicted the heads at several wells in upland areas. It was inferred that these wells intersected, or were near zones of relatively high permeability that were broad enough to extend beneath a nearby stream or lake. This configuration reduced the head in the well, and the hydraulic conductivity of the zone was increased until either the head was reduced sufficiently, or an upper limit of hydraulic conductivity was reached. This was done by including zones of high hydraulic conductivity in the upper bedrock in the vicinity of wells MW-08BR and CW-4 (Figure 7). The heads at MW-06BR and CW-4 were lower than predicted by the model, so a similar approach was used. A high permeability zone was created that extended under the ash basin lake. Slug tests at MW-06BR indicated the hydraulic conductivity of enveloping rock was quite low, so it was assumed the high K zone extended below the depth of the well. This is the motivation for including the high K zone in Layer 12 (Figures 7 and 8). The maximum hydraulic conductivity determined from slug tests in fractured rock was 6 ft/d and this was the upper limit of K assumed to represent these zones in the model. The occurrence of the flat -lying zones of high hydraulic conductivity as shown in Figure 7 and 8 was inferred at the Mayo site based on head observations, but this inference is based on a geologic style that is known from other locations. Flat -lying zones of interconnected fractures several hundred ft or more across were described in crystalline rock at the USGS Mirror Lake research site (e.g. Tiedeman et al. 2001), and similar fracture zones have been recognized at other fractured rock sites that have been studied in detail. It was assumed that the fractures zones in the model were shaped like flat -lying layers several hundred ft in maximum dimension, similar to those described by Tiedeman et al. (2001). The observed hydraulic heads in MW-4 were significantly above the heads predicted by the model calibrated using PEST and adjusted using the fractures zones described above. The head in this well was more than 15 ft above the water level of the nearby ash basin. To account for this difference, a low permeability zone was assumed between MW-4 and the ash basin. This zone was assumed to extend from the saprolite to the sparsely fractured rock at depth, and it was assigned a hydraulic conductivity of 0.01 ft/d. The zone of low permeability used in the model corresponds to a ridge where large boulders of recrystallized quartz were observed in the field. Similar quartz -forming ridges are known elsewhere in the Piedmont where they can form zones of low hydraulic conductivity (e.g. Snipes et al. 1984), presumably because the quartz fills pore space and resists weathering. This zone is intersected by MW-4 and MW-6BR. The water level in both wells changed slowly after it was perturbed, an indication that the wells were completed in tight rocks. The water level change was too slow to obtain meaningful estimates of hydraulic conductivity using slug tests. The flow calibration was done iteratively with transient transport simulations. This was necessary in order to match both the heads and the boron concentration distributions. The fracture zone in Layer 12 plays an important role in the transport. Boron would not reach CW-2 in the absence of a fracture zone, according to the model. As a result, the position of the fracture zone in Layer 12 was adjusted to extend the length of the boron plume to better match the observed data. The final calibrated flow model has the following volume balance: Volume balance in steady state model in ft Feature Input Output Constant Head 59979 113144 Recharge 308190 0 Wells 3150 Drains (streams) 251875 Total 368170 368170 The difference between the input and output is 0.094 ft3, which is a volume balance error of less than 10-6. The major input to the model is from recharge with a lesser amount from constant head boundaries. The constant head boundaries creating input to the model are where ground water is flowing into the model from the boundaries around the periphery. The ash basin lake is also a constant head input. The output is split between groundwater discharging to constant head boundaries and drains. The major constant head sink is Mayo Lake and it accounts for about half the flow going to streams in the model. Less than 1 percent of the water input is removed through wells, according to the model. The final calibrated flow model has a mean head residual of -0.56 ft., and a root mean squared head residual of 5.42 ft. The total span of head measurements ranged over 191 ft, from 365 ft to 556 ft. Using this range to normalize the residual gives a normalized root mean square error of 2.84%. A comparison of the observed and simulated water levels is listed in Table 1, and the observed and simulated levels are cross -plotted in Figure 9. Table 2 lists the best -fit hydraulic parameters from the calibration effort. The calibrated conductivity of the ash is 1 ft/d (Table 2). The calibrated conductivity of the saprolite and transition zones are 3.0 and 1.3 ft/d, respectively. The hydraulic conductivity of the upper fracture rock is 0.05ft/d and it decreases to 0.02 ft/d and 0.007 ft/d with depth. The calibrated values of hydraulic conductivity are consistent with values from the slug tests conducted in the ash, transition zone, and upper fractured rock. No tests were conducted below the upper fractured rock hydrostratigraphic unit, and testing in saprolite was limited to 3 wells. The calibrated model predicts the highest hydraulic head occurs south of the ash basin and the lowest heads occur on Crutchfield Branch and Mayo Creek along the northern edge of the model. The heads in the ash basin are in the midrange between these extremes and this is consistent with the basin resembling a flow -through lake, with higher heads on the west, south and east sides, and lower heads to the north (Figures 10, 11, and 12). 5.2 Flow Model Sensitivity Analysis A parameter sensitivity analysis was performed on the calibrated model by systematically increasing and decreasing the main parameters by factors of either 2 or 0.5 from their calibrated value. Table 4 shows the results of the analysis, expressed in terms of the normalized root mean square error (NRMSE) for each simulation. The baseline NRMSE is 2.84%. The flow model showed the highest degree of sensitivity to the upland recharge and to the hydraulic conductivities of the transition zone and saprolite stratigraphic unit. The saprolite and transition zone were saturated beneath the ash basins and in the vicinity of Crutchfield Branch and this accounts for the calibration sensitivity. The model was only weakly sensitive to the hydraulic conductivities of the ash and the deep rock. There is a weak sensitivity to the hydraulic conductivity of the dams. The heads of wells downgradient of the dams are quite sensitive to the hydraulic conductivity of the dams. However, the hydraulic conductivity of the dams is quite low and changing this value by a factor of 2 has little effect on the overall residual. The sensitivity of the model to the pumping rate of the domestic wells was small, but detectable. This is because the heads in several monitoring wells (e.g. MW-12S, MW-14BR, MW-5BR) can be influenced by pumping from the domestic wells. 5.3 Transport Model Calibration and Sensitivity The transport simulations used a steady state flow model based on current conditions. This assumes that the water level in the ash basin was maintained at approximately 480 ft since the plant opened in 1983. Initial simulations predicted that boron concentrations were absent from MW-3 and CW- 2 and CW-2D. Field observations indicated that boron concentrations were elevated in these wells, so the model was adjusted to match these observations. The adjustments included reducing the Kd to Kd=0.12 mL/gm. The position of a fracture zone inferred to occur in layer 12 was also adjusted by moving it closer to Crutchfield Branch. This affected the concentrations in those wells, although it had little effect on the heads in MW-6BR. 5.4 Result of the Transport Simulation The simulated concentrations reasonably match most of the observed concentrations (Table 5). The concentrations that are an exact match in Table 5 are from wells in the ash basin where the concentrations were set as boundary conditions in the model. Some of the wells where boron was detected are in areas where the predicted concentrations gradients are steep, so small changes in location result in significant changes in concentration. This is one factor that explains the differences between predicted and observed concentrations. The simulations over -estimate the observed arsenic concentrations (Table 6) because of the conservative value of Kd that was used. The primary area of boron groundwater contamination occurs beneath the ash basin dam, and contamination beneath the ash basin on the upgradient end of the basin is also predicted (Figures 14 and 15). The largest plume of boron is predicted to occur beneath the eastern side of the dam. The plume extends under the dam and reaches wells CW-2 and CW-2D, approximately 750 ft along Crutchfield Branch from the edge of the water in the ash basin lake (Figures 14 and 15). The model predicts that the overall footprint of the region above the 2L standard shrinks with depth, but the length of the boron plume along Crutchfield Branch is approximately the same from layers 5 through 12. Observations of boron concentration at CW-2 (804 µg/L) and CW-21) (804 µg/L) suggest that boron concentrations decrease with depth, whereas the model predicts that concentrations should increase with depth. This prediction is a consequence of the geometry of hydraulic conductivity structure inferred during calibration. The high conductivity fracture zone inferred in layer 12, and the low conductivity band have an important influence on transport out of the ash basin. Evidence for this influence is the co -location of the zone of highest concentration, the fracture zone, and the low conductivity band, as shown in Figure 15. The leading edge of the plume as defined by the 2L contour lies between the CW-2 wells and Mayo Lake Rd, according to the simulations. This is important because the property line and compliance boundary are along Mayo Lake Rd. Arsenic concentrations occurred above the 2L level in the saprolite beneath the ash basin. Arsenic concentrations decreased sharply with depth, however, because in the lower half of the transition zone (Layer 7) none of arsenic concentrations exceeded 2L standard. 6.0 PREDICTIVE SIMULATIONS OF CORRECTIVE ACTION Once the flow model was calibrated with regard to water levels, and the simulated concentrations in wells around the ash basin closely matched observed concentrations that exceeded the 2L standards, the model was used to predict contaminant distributions for the next 5, 15, and 30 years. The dates for those simulations are referred to in this report as 2020, 2030, and 2045 respectively. Given the limited information provided to the modeling team about the timing of any actual on -site events, the results provided can be thought of as occurring 5, 15, and 30 years after any activities take place and the actual dates can be modified as needed. Three future conditions were evaluated using the flow and transport model: 41' Corrective Action Plan #1 (CAP 1): No Action 41' Corrective Action Plan #2 (CAP 2): Capping Ash in Place '61' Corrective Action Plan #3 (CAP 3): Complete Ash Removal The distribution of recharge, locations of drains, and distribution of material were modified to represent the different corrective actions. For example, the recharge was modified as shown in Figure 17. The hydraulic head distribution was recalculated and then the transport was simulated for each case. The corrective actions changed the hydraulic head in the vicinity of the ash basin (Figure 18) as the engineered designs interacted with the hydrogeologic conditions. This interaction altered the groundwater flow and the transport of dissolved compounds, as shown in the results of the simulations. 6.1 CAP1 - No Action This method relies on natural attenuation processes to reduce the contaminant concentrations over time. In this scenario, the ash basin is left in place without modification and the assumption is made that current recharge and contaminant loading rates from the ash to the underlying formations are held constant. The model of this scenario includes a distribution of recharge and hydraulic properties as described above. The flow system was assumed to be at steady state with respect to the conditions in 2015. Concentrations in the ash were held constant at the measured concentrations. Boron concentrations delineated by the 2L contour are larger in the simulations from 2045 than they are in 2015, but the differences are relatively small. The most apparent difference is beneath the southwest end of the basin where the 3500 mg/l contour has noticeably expanded from 2015 and 2045 in Layer 7 (e.g. Figure 18b to Figure 33). Another important feature is the leading edge of the boron plume along Crutchfield Branch. The leading edge is on the south side of Mayo Lake Rd in 2015 and 30 years later the leading edge has moved from the south to the north side of the road, according to the simulations. For example, compare the leading edge of the plume in Layer 10 shown in Figure 18c to the same location in Figure 34. The location of the leading edge of the plume in these Figures is important because the property line occurs along Mayo Lake Rd, and the compliance boundary is on or near Mayo Lake Rd. Simulated arsenic concentrations in saprolite beneath the ash increased in 2045 compared to the 2015 values. The distribution of arsenic in the saprolite in 2015 is patchy, and by 2045 the space between many of the patches has filled in (compare Figure 17 to Figure 35). The bottom of the transition zone and underlying rock are less than the 2L standard in 2045, according to the simulations. It is worth pointing out that the Kd value for arsenic was on the low side of the range of lab values and the simulations over -estimated several of the observed arsenic concentrations during calibration. 6.2 CAP2 - Capping Ash in Place The ash basin capping in place method involves placing a low permeable liner over the ash basin to contain the ash and to prevent rainwater infiltration. This scenario assumes that there is no recharge within the ash basin. Initial conditions were applied using the existing concentrations in the formations underneath and within the ash basin as of October 2015. The grid was modified by removing grid blocks to simulate the grading operations described in a design drawing provided by AECOM in October 2015 (Figure 16a). The computational grid was modified to resemble the distribution of ash shown in the design. This included removing grid blocks in the vicinity of the dam to account for grading. Grid blocks above those in the baseline grid were not included in this simulation because it was assumed that material would be unsaturated and would therefore not be active in this simulation. Boundary conditions in the vicinity of the ash basin were modified to account for drainage of the basin, removal of the dam, and grading. The constant head boundary conditions representing the basin were removed. Drain boundary conditions were included to represent drains extending across the graded basin. The drain boundary conditions were set up to provide a flow path for water to exit the system, and to limit the development of impounded conditions upstream from the ash. It was assumed that the dam was removed and the vicinity of the dam was restored to the original hydraulic conductivity, which depends on the hydrostratigraphic unit. Calibration to the ground water heads indicated that the low permeability part of the dam extended at least through layer 7 in the model (through the transition zone). The low permeability zone inferred from the calibration was removed in this simulation. The initial concentrations in this simulation were assumed to be the concentrations simulated to occur in 2015 in the baseline model. Concentration in the ash were allowed to vary. This relaxes the assumption made in the No -Action scenario that the concentrations in the ash basin were held constant. The boron concentrations decreased and the contours changed location from 2015 to 2045. The change is particularly noticeable in the saprolite (Layer 5). Compare Figure 18a to Figure 50, for example. The leading edge of the boron plume recedes in this scenario. By 2045, the leading edge has receded by 100 to 200 ft in the vicinity of Mayo Lake Rd (Figures 50, 51 and 52). Arsenic concentrations increase in the saprolite, but the increase is smaller than for the case of No Action. The arsenic concentration in Layer 7 remains below 2L in this scenario. It is important to keep in mind that the simulation of the No Action scenario outlined above assumed that the concentration in the ash basin remained constant at the current values, whereas the Cap -In -Place scenario described in this section assumed the concentration in the ash started at the observed concentration but could vary with time. This change allows the concentration in the ash and at shallow depths to decrease more rapidly than in the No Action scenario. 6.3 CAP3 -3 Complete Ash Removal Ash basin removal involves removing all ash off -site. The location of the off -site landfill was not considered in this modeling effort. This scenario assumes that the initial concentrations in the formation were the concentrations in the simulations as of October 2015. The upper four layers in the original model were rendered inactive to simulate removal of the ash. "Drains" were added to the model along the axes of drainages exposed when the ash was removed. The other major change is that the recharge within the ash basin boundary was set to the ambient recharge value of 0. 00 18 ft/d (6.5 in/yr). Removing the ash significantly reduces the size and concentration of the boron plume in saprolite and transition zone. For example, the size of the plume in the saprolite in Figure 68 appears to be less than 10 percent of the size of the plume in the No Action case in Figure 32. However, the difference in the size of the plume diminishes with depth and the plume in the bedrock 30 years after the ash was removed is approximately the same size as a comparable time in the No Action case. The concentrations in the plume for the case when the ash has been removed are less than those for the No Action case in the bedrock (Figure 18c and Figure 70). Removing the ash has no effect on the arsenic concentration that can be detected from the contour plots. For example, compare Figure 71 to Figure 17. 7.0 REFERENCES Haven, W. T. 2003. Introduction to the North Carolina Groundwater Recharge Map. Groundwater Circular Number 19. North Carolina Department of Environment and Natural Resources. Division of Water Quality, 8 p. Langley, W.G., J. Daniels, and S. Oza, 2015, Sorption Evaluation of the. Roxboro Steam Electric Plant. Charlotte Department of Civil and Environmental Engineering, report prepared for SynTerra, Legrand, H. 1988. Piedmont and Blue Ridge. Back, W., J. Rosenshein, and P. Seaber, eds. 1988. Hydrogeology: The Geology of North America 0-2: The Decade of North American Geology. Boulder, Colorado: Geological Society of America.. Geological Society of America. P. 201-208. McDonald, M.G. and A.W. Harbaugh, 1988, A Modular Three -Dimensional Finite -Difference Ground -Water Flow Model, U.S. Geological Survey Techniques of Water Resources Investigations, book 6, 586 p. Miller, J.A. 1990. Ground Water Atlas of the U.S. South Carolina and vicinity (HA 730-G). USGS. http://pubs.usgs.gov/ha/ha730/ch g/index.html. Niswonger, R.G.,S. Panday, and I. Motomu, 2011, MODFLOW-NWT, A Newton formulation for MODFLOW-2005, U.S. Geological Survey Techniques and Methods 6-A37, 44-. Snipes, D.S., Burnett, L.L., Wylie, J.A., Sacks, L.A., Heaton, S.B., Dalton, G.A., and Israel, B.A., 1984, Indicators of ground -water quality and yield for a public water supply in rock fracture zones of the piedmont: Clemson, S.C., Water Resources Research Institute Report 115, 80 p SynTerra, 2014, L.V. Roxboro Steam Electric Plant, Semora, NC, Water Supply Well Survey Report of Findings, September 30, 2014. SynTerra, 2015, Comprehensive Site Assessment Report, Roxboro Steam Electric Plant, Semora, NC. September 2, 2015. Tiedeman, C.R. and P.A. Hseih. 2001. Assessing and open hole aquifer test in fractured crystalline rock. Ground Water, v. 39, n.l, p 68-78. Trapp, H. and M.A. Horn. 1997. Ground Water Atlas of the U.S. North Carolina and vicinity (HA 730-L). USGS. http://pubs.usgs.gov/ha/ha730/ch_l/L-text4.html. US EPA, 2015, http://www.epa.gov/watersense/pubs/indoor.html accessed 8/26/15. Watermark Numerical Computing, 2004, PEST Model -Independent Parameter Estimation User Manual: 5th Edition. Zheng, C. and P.P. Wang, 1999, MT3DMS: A Modular Three -Dimensional Multi -Species Model for Simulation of Advection, Dispersion and Chemical Reactions of Contaminants in Groundwater Systems: Documentation and User's Guide, SERDP-99-1, U.S. Army Engineer Research and Development Center, Vicksburg, MS. Tables Table 1. Hydraulic head observed at monitoring wells in June 2015 in the vicinity of the Mayo Plant, heads predicted using the steady state model, and their difference shown as a residual. Head (ft) Well Id Observed Predicted Residual ABMW-01 477.93 482.08 -4.15 ABMW-02 481.56 485.94 -4.38 ABMW-02BR 482.75 485.89 -3.14 ABMW-03 483.02 487.88 -4.86 ABMW-03S 483.04 487.86 -4.82 ABMW-04 485.11 485.05 0.06 ABMW-04D 485.06 484.98 0.08 ABMW-04BR 483.42 485.04 -1.62 MW-03BR 421.83 416.21 5.62 MW-05BR 501.94 501.42 0.52 MW-06BR 450.48 443.00 7.48 MW-07D 444.46 447.45 -2.99 MW-07BR 445.94 447.29 -1.35 MW-08S 437.02 439.47 -2.45 MW-08D 433.53 439.53 -6.00 MW-08BR 432.54 439.56 -7.02 MW-09BR 470.42 463.61 6.81 MW-10BR 499.57 497.07 2.50 MW-11BR 488.05 501.27 -13.22 MW-12S 556.6 548.43 8.17 MW-12D 556.9 548.47 8.43 MW-13BR 495.47 495.33 0.14 MW-14BR 504.61 506.55 -1.94 MW-15BR 402.68 408.64 -5.96 MW-16S 366.92 368.23 -1.31 MW-16D 365.52 368.31 -2.79 MW-16BR 365.41 368.62 -3.21 BG-1 510.26 501.66 8.60 BG-2 511.16 519.08 -7.92 CW-1 472.2 473.55 -1.35 CW-1D 471.93 473.76 -1.83 CW-2 375.29 378.08 -2.79 CW-2D 375.29 384.68 -9.39 CW-3 421.88 415.02 6.86 CW-4 427.7 437.32 -9.62 CW-5 501.53 501.34 0.19 CW-6 450.81 443.55 7.26 MW-2 435.11 431.44 3.67 MW-3 377.32 375.59 1.74 MW-4 496.61 499.21 -2.60 P1 443.77 440.15 3.62 P1A 439.81 440.14 -0.33 P2 413.69 414.28 -0.58 P3 394.41 399.04 -4.63 P3A 394.86 399.05 -4.19 P4 417.02 408.49 8.53 MA 418.84 409.13 9.71 Table 2. Calibrated hydraulic parameters. Unit Kh (ft/d) Kh/Kv Ash, West Basin 1 5 Saprolite 3 1 Transition zone 1.3 1 Upper Rock 0.05 1 fracture zone 4-6 1 Mid Rock 0.02 1 fracture zone 2 1 Rock 0.007 1 Dam 0.05 1 Table 3. Recharge indifferent zones in the model. Recharge zone Flux (ft/d) Regional 0.0018 Mayo Lake 0.0 Ash Basin Lake 0.0001 Ash Basin Dam 0.0001 Mayo Lake Dam 0.0001 Lined Pond 0.0001 Table 4. Flow parameter sensitivity analysis. Results are expressed as model normalized root mean square error (NRMSE) of the simulated and observed heads. NRMSE Unit Kh (ft/d) 0.5 x calibrated Calibrated 2 x calibrated Ash 1 2.90% 2.84% 2.84% Saprolite 3 3.22% 2.84% 3.41% Transition zone 1.3 2.94% 2.84% 3.40% Upper Rock 0.05 2.99% 2.84% 2.91% Mid Rock 0.02 2.84% 2.84% 2.90% Rock 0.007 2.86% 2.84% 2.98% Dams 0.05 3.01% 2.84% 2.85% Upland Recharge 0.0018 5.02% 2.84% 6.29% Domestic wells 50 ft3/d 2.98% 2.84% 3.10% Table 5. Boron concentrations observed in 2015 and simulated at observation wells at Mayo Steam Electric Plant. Boron (mg/L) Well Id Observe Simulated ABMW-01 4 4560 ABMW-02 9 9200 ABMW-02BR 0 1 ABMW-03 1 1940 ABMW-03S 1 395 ABMW-04 4 4930 ABMW-04D 5 1970 ABMW-04BR 0 0 MW-03BR 0 0 MW-05BR 0 0 MW-06BR 0 0 MW-07D 0 0 MW-07BR 0 0 MW-08S 0 0 MW-08D 0 0 MW-08BR 0 0 MW-09BR 0 0 MW-IOBR 0 0 MW-11BR 0 0 MW-12S 0 0 MW-12D 0 0 MW-13BR 0 0 MW-14BR 0 0 MW-15BR 0 0 MW-16S 0 0 MW-16D 0 0 MW-16BR 0 0 BG-1 0 0 BG-2 0 0 CW-1 0 0 CW-ID 0 0 CW-2 8 717 CW-213 3 1315 CW-3 0 0 CW-4 0 0 CW-5 0 0 CW-6 0 0 MW-2 0 7 MW-3 8 1372 MW-4 0 0 Table 6. Arsenic concentrations observed in 2015 and simulated at observation wells at Mayo Steam Electric Plant. Arsenic (µg/I) Observed Simulated ABM W-01 3290 3290 ABM W-02 2491 2491 ABM W-02BR <1 3516 ABM W-03 4232 4232 ABMW-03S 1 3706 ABM W-04 4281 4281 ABM W-04D 140 792 ABMW-04BR 1 518 MW-03BR <1 1020 MW-05BR <1 0 MW-06BR <1 193 MW-07D <1 0 MW-07BR <1 0 MW-08S <1 0 MW-08D <1 0 MW-08BR 3 0 MW-09BR <1 0 MW-10BR <1 0 MW-11BR <1 0 MW-12S <1 0 MW-12D <1 0 MW-13BR <1 0 MW-14BR <1 0 MW-15BR <1 0 MW-16S <1 0 MW-16D <1 0 MW-16BR 2 0 BG-1 <1 0 BG-2 <1 0 CW-1 <1 0 CW-1D <1 0 CW-2 <1 0 CW-2D <1 0 CW-3 <1 0 CW-4 <1 0 CW-5 <1 0 CW-6 <1 0 M W-2 <1 0 M W-3 <1 0 M W-4 <1 0 Figures Figure 1. Map of the Mayo Steam Electric Plant showing ash basin, hydrologic features and boundary of model region. Figure 2. Air photo of the Mayo Steam Electric Plant showing ash basins and hydrologic features and boundary of model region (yellow). Figure 3. Fence diagram of the 3D hydrostratigraphic model (Solids) used to generate hydrostratigraphy. I .. i� � ... a;sllplien11213 • pi1671� yur ql� � Illil 31 �f�1�'� YY I lli' i,tL III ly Inll 7 Ip IIIII illllll llllll � �`YII�1 I �Eu egp or 4[p milt' "I" � �111��161U lj I'}j III I{Iy111Y I Figure 5. Distribution of recharge flux used in the Mayo model. Scale in ft/d. Figure 6. Surface water features and wells in the vicinity of the Mayo Plant. Streams are represented as "drains" and shown in green. Lakes and canals are represented as specified head cells (purple) with a stage equal to the ground surface. The head is specified to be in the middle of the transition zone along the external boundary in upland areas (purple bands around the edge of the model). Water wells are shown as yellow squares. HK '.. �� y 5.4 4.5n::. 3.6Js 2.7- 1.8 0.9 *` 0.0 ft/d 7' d 0 5 r 4 T I ia a� L a y e r 5 L Layer E-7 0, Figure 7. Hydraulic conductivity (ft/d) distributions in grid layers in the Mayo model. Low permeability bands are either dams or tight rock. High permeability zones are inferred flat -lying fracture zones. Figure 8. Zones of different hydraulic conductivity in the vicinity of the ash basin. High hydraulic conductivity in individual layers representing fracture zones (white with orange border), and a zone of low permeability from layers 5 through 12 (blue with white border) in the Mayo model. Mayo 560 520 500 U) 480 -0 460 a� 440 E U 42€3 e1#1#1 ;M all#] 350 400 45U 500 55U COZ Observed Heads (ft) Figure 9. Hydraulic head computed at steady state as a function of the observed head in monitoring wells in the vicinity of the Mayo Plant. The line has a slope of 1:1. Head 580 540 500 460 420 380 340 300 �vTITr IVO 04 / O ! 413 C M _-15BR -1 «1 VR 420 MW-13BR -4 W � -5 Q MW 6WR 09BR Figure 10. Steady state hydraulic head distribution predicted in top of bedrock (layer 8 of the model) in the Mayo model. Colored bars indicate difference between predicted and observed hydraulic head at monitoring wells. Green = less than 4.2 ft (2.5% NRMSE); yellow = 4.2 to 9.5 ft (2.5% to 5% NRMSE); red = greater than 9.5 ft (5% NRMSE). Figure 11. Steady state hydraulic head distribution in predicted in top of bedrock (layer 8 of the model) in the Mayo model. Colored bars indicate difference between predicted and observed hydraulic head at monitoring wells. Green = less than 4.2 ft (2.5% NRMSE); yellow = 4.2 to 9.5 ft (2.5% to 5% NRMSE); red = greater than 9.5 ft (5% NRMSE). Figure 12. Steady state hydraulic head distribution in the vicinity of the ash basin (layer 4 of the Mayo model). Colored bars indicate difference between predicted and observed hydraulic head at monitoring wells. Green = less than 4.2 ft (2.5% NRMSE); yellow = 4.2 to 9.5 ft (2.5% to 5% NRMSE); red = greater than 9.5 ft (5% NRMSE). Figure 13. Concentrations of boron in ash used to specify concentrations in the Mayo model. 2 Laye r 5 Sapralite ` �I Layer 10 Fractured Rack f 5\'i tLayer 12 , r,x Ti Fractured Rock 'i Layer 7 , t Transition zo e. �- ejr Layer 8 Layer 13 Fractured Rock Rock R Ilk,o t Figure 15a. Concentrations of boron predicted to occur in Oct 2015 in Layer 5 of the Mayo model. Red dashed line is outline of inferred fracture zone in Layer 12, and white shaded region with black dashed border is inferred zone of low hydraulic conductivity. Figure 15b. Concentrations of boron predicted to occur in Oct 2015 in Layer 7. Red dashed line is outline of inferred fracture zone in Layer 12, and white shaded region with black dashed border is inferred zone of low hydraulic conductivity. Figure 15c. Concentrations of boron predicted to occur in Oct 2015 in Layer 8. Red dashed line is outline of inferred fracture zone in Layer 12, and white shaded region with black dashed border is inferred zone of low hydraulic conductivity. Figure 15d. Concentrations of boron predicted to occur in Oct 2015 in Layer 10. Red dashed line is outline of inferred fracture zone in Layer 12, and white shaded region with black dashed border is inferred zone of low hydraulic conductivity. Figure 15e. Concentrations of boron predicted to occur in Oct 2015 in Layer 12. Red dashed line is outline of inferred fracture zone in Layer 12, and white shaded region with black dashed border is inferred zone of low hydraulic conductivity. Figure 15f. Concentrations of boron predicted to occur in Oct 2015 in layer 13. Red dashed line is outline of inferred fracture zone in Layer 12, and white shaded region with black dashed border is inferred zone of low hydraulic conductivity. CLOSURE OPTION 2 - CLOSURE IN PLACE PROFILE t11n m am +oa s.ao IQ�00 1'n07 m•aa a3.ma 7MOp 35+00 �WW_" CLOSURE OPTION 2 - CLOSURE IN PLACE CROSS SECTION esa Sao ma 1 Sow W 5q0 10 16 EN00 25 _17 ]0'lY 3B z Figure 16. Map and cross -sections describing design of the Cap -in -Place design used to develop the model of this scenario. Figure 17. Distribution of recharge in simulations of corrective actions. Red is ambient recharge of 0.0018 ft/d, blue is recharge that is negligible. The small region in Layer 5 containing ash is shown where the ash was removed in the lower left panel. Color scale reversed compared to Figure 5. Figure 17a. Hydraulic head calculated in the vicinity of the ash basin for the different corrective actions. Figure 18a. Simulated October, 2015 boron concentrations (µg/L) in the model layer of the saprolite (layer 5). Figure 18b. Simulated October, 2015 boron concentrations (µg/L) in the model layer of the transition zone (layer 7). Figure 18c. Simulated October, 2015 boron concentrations (µg/L) in the model layer of the upper fractured rock (layer 10). Figure 18d. Simulated October, 2015 arsenic concentrations (µg/L) in the model layer of the saprolite (layer 5). Figure 18e. Simulated October, 2015 arsenic concentrations (µg/L) in the model layer of the transition zone (layer 7). Figure 19. Simulated October, 2015 arsenic concentrations (µg/L) in the model layer of the upper fractured rock (layer 10). Figure 20. Simulated October, 2020 boron concentrations (µg/L) in the model layer of the saprolite (layer 5) for CAP I. Figure 21. Simulated October, 2020 boron concentrations (µg/L) in the model layer of the transition zone (layer 7) for CAP 1. Figure 22. Simulated October, 2020 boron concentrations (µg/L) in the model layer of the upper fractured rock (layer 10) for CAP 1. Figure 23. Simulated October, 2020 arsenic concentrations (µg/L) in the model layer of the saprolite (layer 5) for CAP I. Figure 24. Simulated October, 2020 arsenic concentrations (µg/L) in the model layer of the transition zone (layer 7) for CAP 1. Figure 25. Simulated October, 2020 arsenic concentrations (µg/L) in the model layer of the upper fractured rock (layer 10) for CAP 1. Figure 26. Simulated October, 2030 boron concentrations (µg/L) in the model layer of the saprolite (layer 5) for CAP I. Figure 27. Simulated October, 2030 boron concentrations (µg/L) in the model layer of the transition zone (layer 7) for CAP 1. Figure 28. Simulated October, 2030 boron concentrations (µg/L) in the model layer of the upper fractured rock (layer 10) for CAP 1. Figure 29. Simulated October, 2030 arsenic concentrations (µg/L) in the model layer of the saprolite (layer 5) for CAP I. Figure 30. Simulated October, 2030 arsenic concentrations (µg/L) in the model layer of the transition zone (layer 7) for CAP 1. Figure 31. Simulated October, 2030 arsenic concentrations (µg/L) in the model layer of the upper fractured rock (layer 10) for CAP 1. Figure 32. Simulated October, 2045 boron concentrations (µg/L) in the model layer of the saprolite (layer 5) for CAP I. Figure 33. Simulated October, 2045 boron concentrations (µg/L) in the model layer of the transition zone (layer 7) for CAP 1. Figure 34. Simulated October, 2045 boron concentrations (µg/L) in the model layer of the upper fractured rock (layer 10) for CAP 1. Figure 35. Simulated October, 2045 arsenic concentrations (µg/L) in the model layer of the saprolite (layer 5) for CAP I. Figure 36. Simulated October, 2045 arsenic concentrations (µg/L) in the model layer of the transition zone (layer 7) for CAP 1. Figure 37. Simulated October, 2045 arsenic concentrations (µg/L) in the model layer of the upper fractured rock (layer 10) for CAP 1. Figure 38. Simulated October, 2020 boron concentrations (µg/L) in the model layer of the saprolite (layer 5) for CAP2. Figure 39. Simulated October, 2020 boron concentrations (µg/L) in the model layer of the transition zone (layer 7) for CAP2. Figure 40. Simulated October, 2020 boron concentrations (µg/L) in the model layer of the upper fractured rock (layer 10) for CAP2. arsenic : 101812020 12:i]CI�QO;AM 1000.0 100 In + fk-I'a' Figure 41. Simulated October, 2020 arsenic concentrations (µg/L) in the model layer of the saprolite (layer 5) for CAP2. Figure 42. Simulated October, 2020 arsenic concentrations (µg/L) in the model layer of the transition zone (layer 7) for CAP2. Figure 43. Simulated October, 2020 arsenic concentrations (µg/L) in the model layer of the upper fractured rock (layer 10) for CAP2. Figure 44. Simulated October, 2030 boron concentrations (µg/L) in the model layer of the saprolite (layer 5) for CAP2. Figure 45. Simulated October, 2030 boron concentrations (µg/L) in the model layer of the transition zone (layer 7) for CAP2. Figure 46. Simulated October, 2030 boron concentrations (µg/L) in the model layer of the upper fractured rock (layer 10) for CAP2. arsenic : 101612030 1 1000.0 100 o Figure 47. Simulated October, 2030 arsenic concentrations (µg/L) in the model layer of the saprolite (layer 5) for CAP2. Figure 48. Simulated October, 2030 arsenic concentrations (µg/L) in the model layer of the transition zone (layer 7) for CAP2. Figure 49. Simulated October, 2030 arsenic concentrations (µg/L) in the model layer of the upper fractured rock (layer 10) for CAP2. Figure 50. Simulated October, 2045 boron concentrations (µg/L) in the model layer of the saprolite (layer 5) for CAP2. Figure 51. Simulated October, 2045 boron concentrations (µg/L) in the model layer of the transition zone (layer 7) for CAP2. Figure 52. Simulated October, 2045 boron concentrations (µg/L) in the model layer of the upper fractured rock (layer 10) for CAP2. arsenic : 1011012045 12 1000.0 100 Figure 53. Simulated October, 2045 arsenic concentrations (µg/L) in the model layer of the saprolite (layer 5) for CAP2. Figure 54. Simulated October, 2045 arsenic concentrations (µg/L) in the model layer of the transition zone (layer 7) for CAP2. Figure 55. Simulated October, 2045 arsenic concentrations (µg/L) in the model layer of the upper fractured rock (layer 10) for CAP2. Figure 56. Simulated October, 2020 boron concentrations (µg/L) in the model layer of the saprolite (layer 5) for CAP3. AM Figure 57. Simulated October, 2020 boron concentrations (µg/L) in the model layer of the transition zone (layer 7) for CAP3. AM Figure 58. Simulated October, 2020 boron concentrations (µg/L) in the model layer of the upper fractured rock (layer 10) for CAP3. Figure 59. Simulated October, 2020 arsenic concentrations (µg/L) in the model layer of the saprolite (layer 5) for CAP3. ! s �MMEEK c - Figure 60. Simulated October, 2020 arsenic concentrations (µg/L) in the model layer of the transition zone (layer 7) for CA1 3. ! t '41MMEEK- L - Figure 61. Simulated October, 2020 arsenic concentrations (µg/L) in the model layer of the upper fractured rock (layer 10) for CAP3. AM Figure 62. Simulated October, 2030 boron concentrations (µg/L) in the model layer of the saprolite (layer 5) for CAPS. AM Figure 63. Simulated October, 2030 boron concentrations (µg/L) in the model layer of the transition zone (layer 7) for CAP3. r ---- ..-� r ter. - 7-.1- --.`3r Figure 64. Simulated October, 2030 boron concentrations (µg/L) in the model layer of the upper fractured rock (layer 10) for CAP3. 0 � Figure 65. Simulated October, 2030 arsenic concentrations (µg/L) in the model layer of the saprolite (layer 5) for CAPS. 17m� - - Figure 66. Simulated October, 2030 arsenic concentrations (µg/L) in the model layer of the transition zone (layer 7) for CAP3. Figure 67. Simulated October, 2030 arsenic concentrations (µg/L) in the model layer of the upper fractured rock (layer 10) for CAP3. Figure 68. Simulated October, 2045 boron concentrations (µg/L) in the model layer of the saprolite (layer 5) for CAP3. Figure 69. Simulated October, 2045 boron concentrations (µg/L) in the model layer of the transition zone (layer 7) for CAP3. Figure 70. Simulated October, 2045 boron concentrations (µg/L) in the model layer of the upper fractured rock (layer 10) for CAP3. Figure 71. Simulated October, 2045 arsenic concentrations (µg/L) in the model layer of the saprolite (layer 5) for CAP3. Figure 72. Simulated October, 2045 arsenic concentrations (µg/L) in the model layer of the transition zone (layer 7) for CAP3. Figure 73. Simulated October, 2045 arsenic concentrations (µg/L) in the model layer of the upper fractured rock (layer 10) for CAP3. UPDATE SIMULATIONS FOR GROUNDWATER FLOW AND TRANSPORT REPORT FOR MAYO STEAM ELECTRIC PLANT, ROXBORO,NC February 2, 2016 Prepared for SynTerra 148 River Street Greenville, SC 29601 Investigators Lawrence C. Murdoch, Ph.D. Regina Graziano, M.S. Scott E. Brame, M.S. Ronald W. Falta, Ph.D. This report is a continuation of the Groundwater Flow and Transport Modeling Report for Mayo Steam Electric Plant (Murdoch et. al., 2015) [Attachment E in CAP 1 report]. The purpose of this report is to simulate manganese transport at the Mayo Steam Electric Plant for the No Action and the Cap -in -Place scenarios. In addition, the time projection for scenarios No Action and Cap -in -Place is increased from 30 to 100 years for boron transport in groundwater. Simulation of Manganese Transport in Ground Water This section describes the analysis of manganese transport in the vicinity of the ash basin at the Mayo Plant. During summer and fall 2015, a groundwater flow model was developed based on geologic data obtained during drilling of a series of wells in the vicinity of the Plant, and the hydraulic conductivity of the hydrogeologic units was estimated by comparing model predictions of hydraulic heads to values observed in the wells. Transport simulations were conducted using the calibrated flow model to evaluate the migration of boron and arsenic. Concentrations observed in monitoring wells in the ash basin were used as initial conditions, and estimated Kd values were used to estimate the distribution of those compounds in the ground water in the vicinity of the ash basin. After the transport model was calibrated, it was used to estimate the performance of corrective actions, including No Action, Cap -In -Place, Ash Removal, and a Hybrid option. These analyses were described in the Groundwater Flow and Transport Modeling Report for Mayo Steam Electric Plant (Murdoch et. al., 2015). After the original report (Murdoch et. al., 2015) was prepared, it was decided that manganese was of sufficient concern as a constituent to warrant including it in the transport analyses. The procedure was to add manganese as a constituent and simulate transport from 1983 when the basin became operational until 2015. It was included as initial conditions at concentrations ranging from 800 ppb to 2000 ppb in the ash basin based on analyses from monitoring wells. The manganese transport in the Cap -in -Place corrective action scenario was simulated because it is the currently viewed as the most likely scenario to be implemented. The distribution coefficient, Kd, was estimated to vary as a function of pH and the concentrations of other competing compounds, according to analyses by Brian Powell. The average observed pH was 6.5, and this indicates that the average Kd value is 0.1 1/kg, according to Figure 1. The simulation of the Cap -in -Place scenario used for this evaluation was patterned after the scenario described in the CAP 1 report, with several changes. The most notable change is in the conditions used to simulate concentrations in the ash. In the No Action scenario, the concentration distribution from 2015 was used as initial conditions, but other constraints were lifted. In particular, the concentrations in the ash were fixed at the measured concentrations from 1983-2015, but this constraint was lifted in the No Action scenario for simulations of the corrective action. This allowed ambient water to infiltrate the ash and reduce the concentration of the pore water. In this analysis, the constraint that the concentrations in the ash are constant and equal to the observed concentrations was maintained throughout the simulation, from 1983-2045. This approach will cause the concentrations predicted for future corrective actions to exceed the previous approach, which considered reduction of concentrations by dilution in the ash. This change in approach was made to be more conservative. In reality, infiltration will dilute the concentrations in the ash to some extent, so by ignoring this dilution effect the approach used here will tend to give an upper bound on the expected concentrations. Results According to the simulation, manganese is transported downward out of the ash basin, beneath the ash basin dam, and manganese forms a plume extending along Crutchfield Branch (Figs 2-9). The analysis indicates that manganese follows the same flow paths as boron, and indeed, the distribution of manganese predicted by the model is remarkably similar to that of boron. The maximum extent of the plume in 2015 is in the vicinity of the compliance boundary, according to the simulations (Figs 2-9). The simulation is able to approximate some of the observed concentrations in the vicinity of Crutchfield Branch and beneath the ash basin. There are six wells where both the observed and simulated concentrations are above the 2L standard (Table 1). The simulations indicate that some or all of the manganese observed in water from these wells can be explained as having a source from within the ash basin. Concentrations at 19 wells are above the 2L standard and the simulations indicate a concentration of 0. The manganese in the water from these 19 wells cannot be explained as having a source from within the ash basin. Concentrations in water from seven wells are predicted to be below the 2L standard, and this is consistent with observations (Table 1). The simulations indicate that approximately 3/4 of the wells where manganese exceeded the 2L standard represent naturally occurring manganese that is not included in the model. The other '/4 of the affected wells are located in a region influenced by the manganese plume in the simulations. The most striking change in the ground water is a reduction in concentration of manganese following the Cap -in -Place corrective action. This is apparent by comparing the location and extent of the 1100 ppb contours in the same layer at different times (e.g compare Figure 3 to Figure 14, or Figure 4 to Figure 15). This reduction in concentration is also apparent in cross-section (compare Figure 9 to Figure 19). The leading edge of the plume marked by the 2L contour receded slightly 30 years after corrective action (in year 2045 in the simulation). For example, the leading edge was approximately 50 ft beyond the compliance boundary along Crutchfield Branch at 2015 (Fig. 5), but the maximum extent was at or within the compliance boundary in the 2045 simulation (Fig. 16). While encouraging, this change is small and within a reasonable expectation of uncertainty in the location of the leading edge in the simulations. The concentrations of the plume decreased significantly and the plume decreased in size 100 years after the Cap -in -Place corrective action (in year 2115 in the simulation). This simulation predicts the extent of the plume 100 ft away from the compliance boundary, which is farther away from compliance boundary than the 2045 boron simulation prediction. Simulation of Boron Transport in Groundwater —100 year projection A 100 year projection for the No Action and the Cap -in -Place scenarios for boron transport in groundwater were simulated. Originally, the groundwater modeling report (Murdoch et. al., 2015) projected 5, 15, and 30 years. The dates for those simulations are referred to in the groundwater modeling report (Murdoch et. al., 2015) as 2020, 2030, and 2045 respectively. In this report, the boron simulations are projected 100 years after the corrective action plans take place. The date for the simulations is 2115. Details of the No Action and Cap -in -Place modeling method can be found in the Groundwater Flow and Transport Modeling Report for Mayo Steam Electric Plant (Murdoch et. al., 2015). No Action Figures 47 through 49 depict the transport of boron for the No Action scenarios within model layers 5 (saprolite), 7 (transition), and 10 (upper fractured bedrock). The simulated 2115 boron concentrations delineated by the 2L contour within the saprolite is similar to the 2045 simulation (Murdoch et. al., 2015). The 2115 boron plumes within the transition zone and bedrock are slightly bigger than the 2045 simulations. The simulated 2115 boron bedrock plume has two locations where the edge of the plume has migrated slightly north of Mayo Lake Road, which is the property and compliance boundary. The leading edge of the 2115 boron plume within the saprolite and transition zone are just south of Mayo Lake Road, which is similar to the 2045 simulations. Cap -in -Place Figures 50 through 53 depict the transport of boron for the Cap -in -Place scenarios within model layers 5, 7, and 10. The boron concentrations decreases from 2045 to 2115 and is noticeable in the saprolite, transition zone, and fractured bedrock layers. The leading edge of the boron plume also recedes from 2045 to 2115. By 2115, the leading edge in some areas has receded by 200 to 300 ft in the vicinity of Mayo Lake Road. References Langley, W.G., J. Daniels, and S. Oza, 2015, Sorption Evaluation of the. Roxboro Steam Electric Plant. Charlotte Department of Civil and Environmental Engineering, report prepared for SynTerra, McDonald, M.G. and A.W. Harbaugh, 1988, A Modular Three -Dimensional Finite -Difference Ground - Water Flow Model, U.S. Geological Survey Techniques of Water Resources Investigations, book 6, 586 p. Murdoch, L.C., S. E. Brame, R. W. Falta, and R. A.Graziano, November 2015. Groundwater Flow and Transport Modeling Report for Mayo Steam Electric Plant, Roxboro, NC. Niswonger, R.G.,S. Panday, and I. Motomu, 2011, MODFLOW-NWT, A Newton formulation for MODFLOW-2005, U.S. Geological Survey Techniques and Methods 6-A37, 44-. SynTerra, 2015, Comprehensive Site Assessment Report, Roxboro Steam Electric Plant, Semora, NC. September 2, 2015. Powell, Brian, November 2015. Analysis of Geochemical Phenomena Controlling Mobility of Ions from Coal Ash Basins at the Duke Energy Mayo Steam Electric Plant. Pendleton, SC. Zheng, C. and P.P. Wang, 1999, MT3DMS: A Modular Three -Dimensional Multi -Species Model for Simulation of Advection, Dispersion and Chemical Reactions of Contaminants in Groundwater Systems: Documentation and User's Guide, SERDP-99-1, U.S. Army Engineer Research and Development Center, Vicksburg, MS. Mn Kd values versus pH in three G\%* silnulants 100E-01 1.00E-00 100E41 100E-0' 1 3 S 6 11 6 9 10 1H *%%Kd. \SIB (', t' Vxhxs •Mh Kd. AVGGn' Fives & %% Kd. NLAA GR' Fives Figure 1. Distribution coefficient for manganese as a function of pH and relative concentration of competing ions. Figure from report by Brian Powell. Table 1. Concentrations of manganese observed in monitoring wells and simulated using the model. Rows shaded brown were fixed concentrations in the model. The 6 rows shaded yellow indicate concentrations that are above 2L in both the field observations and the simulations. The 19 rows shaded green indicate field observations that are above 2L, but the simulated values are less than 2L. The 7 rows shaded pink are where both the observations and the simulations are less than 2L. observed simulated well name Mn Mn A B M W -01 2000 2000 A B M W -02 1000 1000 ABMW-02BR 149 1 A B M W -03 1000 1000 ABMW-03S 478 203 A B M W -04 800 800 ABMW-04D 3640 337 ABMW-04BR 407 0 MW-03BR 731 0 MW-OSBR 1740 0 MW-07D 188 0 MW-07BR 263 0 MW-08BR 5040 0 MW-09BR 809 0 MW-10BR 894 0 MW-11BR 67 0 MW-12S 267 0 MW-12D 683 0 MW-13BR 474 0 MW-14BR 544 0 MW-16S 260 0 MW-16D 269 0 MW-16BR 30 0 BG-1 35 0 BG-2 150 0 CW-1 27 0 CW-1D 6 0 CW-2 170 250 CW-2D 123 394 CW-3 9 0 CW-4 0 0 CW-5 696 0 CW-6 1300 0 MW-2 833 7 M W -3 880 400 M W -4 15 0 Figure 2. Simulated 2015 manganese concentrations (µg/L) in model layer 4 (ash). Figure 3. Simulated 2015 manganese concentrations (µg/L) in model layer 5 (saprolite). Figure 4. Simulated 2015 manganese concentrations (µg/L) in model layer 7 (transition zone). Figure 5. Simulated 2015 manganese concentrations (µg/L) in model layer 9 (upper rock). Figure 6. Simulated 2015 manganese concentrations (µg/L) in model layer 11 (rock). Figure 8. Simulated 2015 manganese concentrations (µg/L) in model layer 13 (rock). manganese: 10/10/2015 12:01:24AM Figure 9. Simulated 2015 manganese concentrations (µg/L) in cross-section trending along Crutchfield Branch. Purple line shows top of rock. Figure 10. Simulated 2045 manganese concentrations (µg/L) in model layer 1 (ash) for No Action. Figure 11. Simulated 2045 manganese concentrations (µg/L) in model layer 2 (ash) for No Action. Figure 12. Simulated 2045 manganese concentrations (µg/L) in model layer 3 (ash) for No Action. Figure 13. Simulated 2045 manganese concentrations (µg/L) in model layer 4 (ash) for No Action. Figure 14. Simulated 2045 manganese concentrations (µg/L) in model layer 5 (saprolite) for No Action. Figure 15. Simulated 2045 manganese concentrations (µg/L) in model layer 7 (transition zone) for No Action. Figure 16. Simulated 2045 manganese concentrations (µg/L) in model layer 9 (shallow rock) for No Action. Figure 17. Simulated 2045 manganese concentrations (µg/L) in model layer 11 (rock) for No Action. Figure 18. Simulated 2045 manganese concentrations (µg/L) in model layer 13 (rock) for No Action. Figure 19. Simulated 2115 manganese concentrations (µg/L) in model layer 1 (ash) for No Action. Figure 20. Simulated 2115 manganese concentrations (µg/L) in model layer 2 (ash) for No Action. Figure 21. Simulated 2115 manganese concentrations (µg/L) in model layer 3 (ash) for No Action. Figure 22. Simulated 2115 manganese concentrations (µg/L) in model layer 4 (ash) for No Action. Figure 23. Simulated 2115 manganese concentrations (µg/L) in model layer 5 (saprolite) for No Action. Figure 24. Simulated 2115 manganese concentrations (µg/L) in model layer 7 (transition zone) for No Action. Figure 25. Simulated 2115 manganese concentrations (µg/L) in model layer 9 (shallow rock) for No Action. Figure 26. Simulated 2115 manganese concentrations (µg/L) in model layer 11 (rock) for No Action. Figure 27. Simulated 2115 manganese concentrations (µg/L) in model layer 13 (rock). Assumes Cap -in -Place corrective action. Figure 28. Simulated 2045 manganese concentrations (µg/L) in model layer 1 (ash). Assumes Cap -in -Place corrective action. Figure 29. Simulated 2045 manganese concentrations (µg/L) in model layer 2 (ash). Assumes Cap -in -Place corrective action. Figure 30. Simulated 2045 manganese concentrations (µg/L) in model layer 3 (ash). Assumes Cap -in -Place corrective action. Figure 31. Simulated 2045 manganese concentrations (µg/L) in model layer 4 (ash). Assumes Cap -in -Place corrective action. Figure 32. Simulated 2045 manganese concentrations (µg/L) in model layer 5 (saprolite). Assumes Cap -in -Place corrective action. Figure 33. Simulated 2045 manganese concentrations (µg/L) in model layer 7 (transition zone). Assumes Cap -in -Place corrective action. Figure 34. Simulated 2045 manganese concentrations (µg/L) in model layer 9 (shallow rock). Assumes Cap -in -Place corrective action. Figure 35. Simulated 2045 manganese concentrations (µg/L) in model layer 11 (rock). Assumes Cap -in -Place corrective action. Figure 36. Simulated 2045 manganese concentrations (µg/L) in model layer 13 (rock). Assumes Cap -in -Place corrective action. manganese : M9I2045 11:58:36 PM Figure 37. Simulated 2045 manganese concentrations (µg/L) in cross-section trending along Crutchfield Branch. Assuming Cap -in -Place corrective action. Figure 38. Simulated 2115 manganese concentrations (µg/L) in model layer 1 (ash). Assumes Cap-m- Place corrective action. Figure 39. Simulated October, 2115 manganese concentrations (µg/L) in model layer 2 (ash). Assumes Cap -in -Place corrective action. Figure 40. Simulated October, 2115 manganese concentrations (µg/L) in model layer 3 (ash). Assumes Cap -in -Place corrective action. Figure 41. Simulated October, 2115 manganese concentrations (µg/L) in model layer 4 (ash). Assumes Cap -in -Place corrective action. Figure 42. Simulated October, 2115 manganese concentrations (µg/L) in model layer 5 (saprolite). Assumes Cap -in -Place corrective action. Figure 43. Simulated October, 2115 manganese concentrations (µg/L) in model layer 7 (transition zone). Assumes Cap -in -Place corrective action. Figure 44. Simulated October, 2115 manganese concentrations (µg/L) in model layer 9 (shallow rock). Assumes Cap -in -Place corrective action. Figure 45. Simulated October, 2115 manganese concentrations (µg fl in model layer 11 (rock). Assumes Cap -in -Place corrective action. Figure 46. Simulated October, 2115 manganese concentrations (µg/L) in model layer 13 (rock). Assumes Cap -in -Place corrective action. Figure 47. Simulated October, 2115 boron concentrations (µg/L) in the model layer of the saprolite (layer 5) for No Action. Figure 48. Simulated October, 2115 boron concentrations (µg/L) in the model layer of the transition zone (layer 7) for No Action. Figure 49. Simulated October, 2115 boron concentrations (µg/L) in the model layer of the upper fractured rock (layer 10) for No Action. Figure 50. Simulated October, 2115 boron concentrations (µg/L) in the model layer of the saprolite (layer 5). Assumes Cap -in -Place corrective action. Figure 51. Simulated 2115 boron concentrations (µg/L) in the model layer of the transition zone (layer 7). Assumes Cap -in -Place corrective action. Figure 52. Simulated 2115 boron concentrations (µg/L) in the model layer of the upper fractured rock (layer 10). Assumes Cap -in -Place corrective action.