Loading...
HomeMy WebLinkAboutNC0038377_CSA October 2017 Report_201711012017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 2 of 2 Explanation of altered or items not initialed: Item 1. The CSA was specifically designed to assess the coal ash management areas of the facility. Sufficient information is available to prepare the groundwater corrective action plan for the ash management areas of the facility. Data limitations are discussed in Section 11.0 of the CSA report. Continued groundwater monitoring at the Site is planned. Item 2. Imminent hazards to human health and the environment have been evaluated. The NCDEQ data associated with nearby water supply wells is provided herein and is being evaluated. Item 5. The groundwater assessment plan for the CSA as approved by NCDEQ was specifically developed to assess the coal ash management areas of the facility for the purposes of developing a corrective action plan for groundwater. Other areas of possible contamination on the property, if noted, are anticipated to be evaluated separately. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page ES-1 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx EXECUTIVE SUMMARY ES.1 Source Information Duke Energy Progress, LLC (Duke Energy) owns and operates the Mayo Steam Electric Plant (the Mayo Plant, Plant, or Site), located in Person County, near Roxboro, North Carolina. The Comprehensive Site Assessment (CSA) update was conducted to refine and expand the understanding of subsurface conditions and evaluate the extent of impacts from historical management of coal ash. This CSA update contains an assessment of site conditions based on a comprehensive interpretation of geologic and sampling results from the initial site assessment and geologic and sampling results obtained subsequent to the initial assessment. The Mayo Plant began operations in 1983 and is presently in service. At the Site, there is a single ash basin located northwest of the Plant. The ash basin, constructed with an earthen dike, is approximately 140 acres and contains ash generated from the Plant’s historic coal combustion. Discharge from the ash basin via Outfall 002 to Mayo Lake is permitted by the NCDEQ Division of Water Resources (NCDEQ-DWR) under National Pollutant Discharge Elimination System (NPDES) Permit NC0038377. A lined coal combustion residuals (CCR) landfill is located to the west of the Plant on the west side of US Highway 501 (Boston Road). No other areas of coal ash are known to exist at the Site. CCR was managed at the Plant’s on-site ash basin and transported via wet sluicing until 2013 when the Mayo Plant converted to a dry ash system in which 90 percent of generated CCR was dry. After the conversion, wet sluicing was conducted only when there was a shutdown of the dry fly ash transport system. Final system upgrades were completed in October 2016 and all CCR collection is now dry. Prior to November 2014, dry ash was hauled to and disposed of in the lined landfill located at the nearby Roxboro Steam Electric Plant (near Semora, North Carolina). Since November 2014, CCR from the Mayo Plant has been placed in a newly constructed on-site Industrial landfill (monofill; Permit #7305) at the Mayo Site. Assessment results indicate the thickness of CCR in the portion of the ash basin not covered with water is approximately 13 to 66 feet. Assessment findings determined that CCR accumulated in the ash basin is the source of impact to groundwater. The inferred general extent of constituent migration from the ash basin based on evaluation of constituent concentrations greater than both water quality standards and background is shown on Figure ES-1. A detailed evaluation of constituent migration is included in the CSA update report. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page ES-2 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx ES.2 Initial Abatement and Emergency Response Duke Energy has not conducted emergency responses because groundwater impacts from the ash basin do not present an imminent and substantial threat to the environment requiring emergency action. Regarding initial abatement, as previously described, upgrades to the dry fly ash handling system were completed in October 2016 and all CCR collection is now dry. In preparation for ash basin closure, new retention basins and wastewater treatment systems are being designed and constructed. ES.3 Receptor Information In accordance with North Carolina Department of Environmental Quality (NCDEQ) direction, CSA receptor survey activities include listing and depicting all water supply wells (public or private, including irrigation wells and unused wells) within a 0.5-mile radius of the ash basin compliance boundary. ES.3.1 Public Water Supply Wells According to public records, Bethel Hill Baptist Church, located approximately 0.5 miles south (upgradient) of the Plant at 201 Old US Highway maintains a public water supply provided by a groundwater well. ES.3.2 Private Water Supply Wells No private water supply wells are located within 0.5 mile downgradient of the ash basin compliance boundary. Drinking water is obtained from approximately 20 private groundwater wells by residences within a 0.5-mile radius of the Mayo Plant compliance boundary and located to the south and northwest the Plant, upgradient of the predominant groundwater flow direction and upgradient of the ash basin. In 2015, NCDEQ coordinated sampling of three of the private water supply wells, and in 2017, Duke Energy collected samples from eight additional private water supply wells. Available analytical data for the private wells generally show detected concentrations below statistically derived background concentrations and geochemistry (cation/anion distribution) not attributable to CCR impacts. The supply well data should be interpreted cautiously because well construction and equipment (e.g., galvanized piping, pump components) may influence analytical results. The water supply wells are, without exception, located upgradient of the Mayo Plant ash basin, and modeling has demonstrated that well capture zones are limited to the immediate vicinity of the well head and do not extend toward the ash basin. The land directly downgradient of the ash basin and the Duke property line is undeveloped; therefore, no water supply wells are located north and 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page ES-3 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx downgradient of the groundwater plume within the survey radius. The updated Corrective Action Plan (CAP) will address any potential future risk to future groundwater receptors. ES.3.3 Surface Water Bodies The Site is located in the Roanoke River Basin. Although the Site is located near Mayo Lake, groundwater influenced by the ash basin primarily flows toward and discharges to the Crutchfield Branch stream valley, a small stream located north of the ash basin and near the northern property boundary. There is no surface water intake located in Crutchfield Branch. ES.3.4 Human and Ecological Receptors A baseline human health and ecological risk assessment was performed in 2016 as a component of the CAP, Part 2 (SynTerra, 2016a), concluding that no unacceptable risks to humans resulted from hypothetical exposure to constituents detected in the ash basin and south creek (upstream) areas. Based on review and analysis of groundwater and surface water data collected since completing the human health and ecological risk assessment in 2016, there is no evidence of potential risks to humans and wildlife at the Mayo Site. This update to the human health and ecological risk assessment supports a risk classification of “Low”. ES.3.5 Land Use Land use surrounding the Mayo Plant includes rural, rural residential, agricultural, and forest land. Duke Energy-owned and maintained land borders the Mayo Plant to the west, east (Mayo Lake), and north (with the exception of two undeveloped parcels). A small residential area borders the Site to the south- southwest. Property within 500 feet of the Mayo Plant compliance boundary is owned by Duke Energy with the exception of the undeveloped parcels located due north of the northern Plant boundary along Mayo Lake Road. No change in land use surrounding the Mayo Site is currently anticipated. ES.4 Sampling/Investigation Results The comprehensive site assessment included evaluations of the hydrogeological and geochemical properties of soil and groundwater at multiple depths and distances from the ash basin. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page ES-4 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx ES.4.1 Background Concentration Determinations Naturally occurring background concentrations were determined using statistical analysis for both soil and groundwater. Statistical determinations of provisional background threshold values (PBTVs) were performed in strict accordance with the revised Statistical Methods for Developing Reference Background Concentrations for Groundwater and Soil at Coal Ash Facilities (statistical methods document) (HDR and SynTerra, 2017). The current background monitoring well network consists of wells installed within three flow zones – surficial, transition zone, and fractured bedrock. Background datasets for each flow system used to statistically determine naturally occurring concentrations of inorganic constituents in soil and groundwater are provided herein. As of September 1, 2017, NCDEQ approved a number of the statistically derived background values; however, others are still under evaluation and thus considered preliminary at this time. Background results may be greater than the PBTVs due to the limited valid dataset currently available. The statistically derived background threshold values will continue to be adjusted as additional data becomes available. ES.4.2 Nature and Extent of Contamination Site-specific groundwater constituents of interest (COIs) were developed by evaluating groundwater sampling results with respect to 2L/IMAC and PBTVs, and additional regulatory input/requirements. The distribution of constituents in relation to the ash basin, co-occurrence with CCR indicator constituents such as boron, and likely migration directions based on groundwater flow direction are considered in determination of groundwater COIs. The following list of groundwater COIs has been developed for Mayo: Arsenic Manganese Barium Molybdenum Boron pH Chromium (hexavalent) Strontium Chromium (total) Sulfate Cobalt Total Dissolved Solids (TDS) Iron Vanadium Wells monitoring the surficial, transition zone, and bedrock flow units were installed beneath the ash basin. Wells completed in the saprolite or transition zone beneath the ash basin have PBTV and 2L exceedances for arsenic, barium, 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page ES-5 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx boron, cobalt, iron, manganese, molybdenum, strontium, TDS and vanadium (a number of which only occur in the transition zone). For the most recent monitoring event (March-April 2017), bedrock monitoring wells installed within the ash basin indicate only strontium is detected greater than the PBTV. There is no 2L standard for strontium. The remaining constituents were not detected in the bedrock below the ash basin at concentrations greater than PBTV and 2L. At Mayo, boron, manganese, and strontium are key indicators of CCR groundwater impacts. Boron is detected at concentrations greater than the 2L standard beneath and downgradient (north-northeast) of the ash basin. Boron is not detected in background groundwater. Manganese and strontium are detected at concentrations greater than PBTV. The area farthest downgradient at which boron, manganese, and strontium are detected at a concentration greater than applicable 2L and PBTVs is interpreted as the leading edge of the CCR-derived plume moving downgradient from the source area. For the surficial flow unit, boron, manganese, and strontium were detected in monitoring wells screened in Crutchfield Branch alluvium downgradient of the ash basin. In the transition zone and bedrock flow units, boron and strontium are detected above PBTVs in downgradient wells closest to the northern property line. Manganese was detected above PBTVs in a bedrock well near the compliance boundary. Boron was not detected in groundwater from wells located approximately 1,000 feet downgradient of the property. The leading edge of the transition zone and bedrock plume is interpreted to be at/near the northern property line. The surficial and transition zone flow units at Mayo, though impacted, are not vertically extensive. Impact to the bedrock flow unit is present in the upper 50 to 75 feet of fractured bedrock. The vertical extent of the plume is represented by groundwater concentrations in bedrock wells beneath and downgradient of the ash basin. ES.4.3 Maximum Contaminant Concentrations (Source Information) The source area at Mayo Plant includes CCR material and pore water accumulated in the ash basin. Ash pore water samples collected from wells installed within the ash basins and screened in the ash layers have been monitored since 2015. The concentrations of detected constituents have been relatively stable with minor fluctuations. The ash basin is a permitted 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page ES-6 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx wastewater system; therefore, comparison of pore water within the wastewater treatment residuals (ash) to 2B or 2L/IMAC is not required. Soil samples collected below the ash/soil interface from three locations within the ash basin indicate only arsenic concentrations in one soil sample both exceeded the calculated soil PBTV and the NCDEQ Preliminary Soil Remediation Goals (PSRG) Protection of Groundwater (POG) value. Several other constituents exceeded the PBTV only. ES.4.4 Site Geology and Hydrogeology The subsurface at the Mayo Site is comprised of a surficial unit (soil, fill and reworked soil, alluvium, and saprolite), a transition zone, and fractured bedrock. The transition zone is comprised of partially weathered rock that is gradational between saprolite and competent bedrock. The bedrock is dominantly granitoid gneiss and mica gneiss with minor mica schist and phyllite. Shallow bedrock is fractured; however, only mildly productive fractures (providing water to wells) were observed within the top 50 – 75 feet of bedrock. The majority of fractures are relatively small (e.g., close and tight) and appear to be limited in connectivity between borings. Yields from pumping or packer testing are low. Groundwater exists under unconfined or water table conditions throughout the Site. For the most part, saturated conditions are limited to secondary fractures within the underlying bedrock. Saturated conditions in saprolite and the transition zone are limited and sporadic across the Site. The hydrogeologic characteristics of the ash basin environment are the primary control mechanisms on groundwater flow and constituent transport. The basin acts as a bowl-like feature toward which groundwater flows from the northwest, west, south, and east. Groundwater flows from the highest topographic portion of the Site (near the Plant entrance road) to the north and northeast. The flow of ponded water within the ash basin is controlled laterally by groundwater flow that enters the basin and is controlled downgradient (north-northeast) by the ash basin dam and the National Pollutant Discharge Elimination System (NPDES) outfall/discharge (east side of basin). The head created by the impounded water in the ash basin creates a slight mounding effect that influences the groundwater flow direction in the immediate vicinity of the basin. East of the ash basin, there is a groundwater divide that separates the Crutchfield Branch flow regime from the Mayo Lake flow regime. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page ES-7 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx The stream valley in which the ash basin was constructed is a distinct slope- aquifer system in which flow of groundwater into the ash basin and out of the ash basin is restricted to the local flow regime. Groundwater and surface water flow from the ash basin is funneled into the small valley formed by Crutchfield Branch and flows north-northeast from the ash basin into the Crutchfield Branch valley, which flows north of the property and into Mayo Creek. ES.5 Conclusions and Recommendations The investigation described in the CSA presents the results of the assessments required by the Coal Ash Management Act (CAMA) and 2L. The ash basin CCR material pore water was determined to be a source of the groundwater contamination. The assessment investigated the Site hydrogeology, determined the direction of groundwater flow from the ash basin, and determined the horizontal and vertical extent of impacts to groundwater and soil sufficient to proceed with preparation of a CAP. Only one soil sample from below the ash basin contained arsenic at a concentration greater than PBTV and the PSRG POG value. Strontium was also detected in the same soil sample at concentrations greater than the PBTV. Arsenic is not detected in groundwater at concentrations greater than the 2L standard beyond the ash basin waste boundary. There is no PSRG POG for strontium. Strontium is present above the groundwater PBTVs beyond the compliance boundary in the surficial aquifer only. No other COIs were detected in soil beneath the ash basin at concentrations greater than both the PBTV and POG. Shallow soil impacts are anticipated to be addressed through basin closure and the CAP. Boron is the primary constituent detected in groundwater at concentrations greater than background and the 2L standard near or beyond the compliance boundary. Manganese and strontium above the respective PBTVs also appear to be indicators of impact to groundwater. The interpreted extent of boron concentrations greater than the 2L standard is near the compliance boundary in the surficial and transition flow zones. The boron concentration is less than 2L standard in the bedrock flow unit near the compliance boundary. The interpreted extent of manganese concentrations greater than the PBTV and 2L standard is beyond the compliance boundary in the surficial and bedrock flow zones; however, the manganese concentration is less than the PBTV within the transition zone at the compliance boundary. The interpreted extent of strontium concentrations greater than the PBTV extends beyond the compliance boundary only within the surficial flow zone. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page ES-8 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx Mayo Plant’s ash basin is currently designated as “Intermediate” risk under CAMA, meaning that closure of the ash basin is required by 2024. The updated evaluation of risks has determined no imminent risk to human health or the environment due to groundwater, surface water, or sediment impacts. The private water supply wells located near the plant are not located in the ash basin groundwater flow path. A "Low" risk classification and closure via a cap-in-place scenario are considered viable. A preliminary evaluation of groundwater corrective action alternatives is included in this CSA to provide insight into the CAP preparation process. For Mayo, the primary source control (closure) methods anticipated to be evaluated in the CAP are: Dewater the ash within the basin and cap the residuals with a low permeability engineered cover system to minimize infiltration; Excavate the ash to remove the source of the COIs from the groundwater flow system; and Potentially some combination of the above. The source control (closure) options will be evaluated in the CAP to determine the most technically and economically feasible means of removing or controlling the ash and ash pore water as a source to the groundwater flow system. The evaluation will include predictive groundwater modeling to evaluate the cost-benefit associated with various options. For basin closure, preliminary modeling indicates ash dewatering and reduction of the amount of water migrating from the basin to groundwater will have the greatest positive impact on groundwater and surface water quality downgradient of the ash basin. A well-designed capping system can be expected to minimize ongoing migration to groundwater after dewatering. In addition to source control measures, the CAP will evaluate measures to address groundwater conditions associated with the ash basin. Groundwater corrective action by monitored natural attenuation (MNA) is anticipated to be a remedy further evaluated in the CAP. As warranted, a number of viable groundwater remediation technologies such as phytoremediation, groundwater extraction, or hydraulic barriers may be evaluated based upon short-term and long-term effectiveness, implementability, and cost. Results of the evaluation, including groundwater fate and transport modeling, and geochemical modeling, will be used for remedy selection in the CAP. 148 RIVER STREET, SUITE 220 GREENVILLE, SOUTH CAROLINA 29601 PHONE 864-421-9999 www.synterracorp.com PROJECT MANAGER: LAYOUT: DRAWN BY: JERRY WYLIE DATE:ADAM FEIGL ES1 - MAYO 10/09/2017 10/29/2017 2:37 PM P:\Duke Energy Progress.1026\00 GIS BASE DATA\Mayo\Map_Docs\CSA_Supplement_2\3D SCM\Mayo_3D_ES1.dwg FIGURE ES-1 APPROXIMATE EXTENT OF IMPACTS MAYO STEAM ELECTRIC PLANT DUKE ENERGY PROGRESS, LLC ROXBORO, NORTH CAROLINA PROGRESS MAYO LAKE MAYO LAKE ROAD BOSTON RD. (HIGHWAY 501)1981 C & D LANDFILL MAYO STEAM ELECTRIC PLANT SURFICIAL TRANSITION ZONE VISUAL AID ONLY - DEPICTION NOT TO SCALE NORTH BEDROCK CRUTCHFIELD BRANCH ASH BASIN ASH BASIN WASTE BOUNDARY GENERALIZED GROUNDWATER FLOW DIRECTION APPROXIMATE LANDFILL WASTE BOUNDARY NOTE: 1.OCTOBER, 2016 AERIAL PHOTOGRAPHY OBTAINED FROM GOOGLE EARTH PRO ON SEPTEMBER 27, 2017, DATED JUNE 13, 2016. 2.STREAM FROM WSP SURVEY, 2014. 3.GENERALIZED GROUNDWATER FLOW DIRECTION BASED ON APRIL 10, 2017 WATER LEVEL DATA. 4.PROPERTY BOUNDARY PROVIDED BY DUKE ENERGY. 5.GENERALIZED AREAL EXTENT OF MIGRATION REPRESENTED BY NCAC 02L EXCEEDANCES. STREAM WITH FLOW DIRECTION RESIDENTIAL UNIT DUKE ENERGY PROPERTY BOUNDARY LEGEND AREA OF CONCENTRATION IN GROUNDWATER ABOVE NC2L (SEE NOTE 5) INFERRED AREA OF CONCENTRATION IN GROUNDWATER ABOVE NC2L (SEE NOTE 5) 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page i P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx TABLE OF CONTENTS SECTION PAGE CERTIFICATION PAGE ES.1 SOURCE INFORMATION ....................................................................................... ES-1 ES.2 INITIAL ABATEMENT AND EMERGENCY RESPONSE ................................ ES-2 ES.3 RECEPTOR INFORMATION .................................................................................. ES-2 ES.3.1 Public Water Supply Wells ................................................................................... ES-2 ES.3.2 Private Water Supply Wells ................................................................................. ES-2 ES.3.3 Surface Water Bodies ............................................................................................. ES-3 ES.3.4 Human and Ecological Receptors ........................................................................ ES-3 ES.3.5 Land Use .................................................................................................................. ES-3 ES.4 SAMPLING/INVESTIGATION RESULTS .......................................................... ES-3 ES.4.1 Background Concentration Determinations ...................................................... ES-4 ES.4.2 Nature and Extent of Contamination .................................................................. ES-4 ES.4.3 Maximum Contaminant Concentrations (Source Information) ...................... ES-5 ES.4.4 Site Geology and Hydrogeology ......................................................................... ES-6 ES.5 CONCLUSIONS AND RECOMMENDATIONS ................................................ ES-7 1.0 INTRODUCTION ......................................................................................................... 1-1 1.1 Purpose of Comprehensive Site Assessment ........................................................ 1-1 1.2 Regulatory Background ........................................................................................... 1-2 Notice of Regulatory Requirements (NORR) ............................................... 1-2 1.2.1 Coal Ash Management Act Requirements .................................................... 1-3 1.2.2 1.3 Approach to Comprehensive Site Assessment ..................................................... 1-4 NORR Guidance ................................................................................................ 1-4 1.3.1 USEPA Monitored Natural Attenuation Tiered Approach ........................ 1-5 1.3.2 ASTM Conceptual Site Model Guidance ....................................................... 1-5 1.3.3 1.4 Technical Objectives ................................................................................................. 1-5 1.5 Previous Submittals .................................................................................................. 1-6 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page ii P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx TABLE OF CONTENTS SECTION PAGE 2.0 SITE HISTORY AND DESCRIPTION ..................................................................... 2-1 2.1 Site Description, Ownership and Use History...................................................... 2-1 2.2 Geographic Setting, Surrounding Land Use, Surface Water Classification ..... 2-2 2.3 CAMA-related Source Areas ................................................................................... 2-5 2.4 Other Primary and Secondary Sources .................................................................. 2-6 2.5 Summary of Permitted Activities ........................................................................... 2-6 2.6 History of Site Groundwater Monitoring .............................................................. 2-8 Ash Basin Voluntary Groundwater Monitoring .......................................... 2-8 2.6.1 Ash Basin NPDES Groundwater Monitoring ............................................... 2-8 2.6.2 Ash Basin CAMA Groundwater Monitoring................................................ 2-9 2.6.3 Landfill Groundwater Monitoring ................................................................. 2-9 2.6.4 2.7 Summary of Assessment Activities ........................................................................ 2-9 2.8 Summary of Initial Abatement, Source Removal or Other Corrective Action .......................................................................................... 2-9 3.0 SOURCE CHARACTERISTICS ................................................................................. 3-1 3.1 Coal Combustion and Ash Handling System ....................................................... 3-1 3.2 General Physical and Chemical Properties of Ash............................................... 3-2 3.3 Site-Specific Coal Ash Data ..................................................................................... 3-4 4.0 RECEPTOR INFORMATION ..................................................................................... 4-1 4.1 Summary of Receptor Survey Activities................................................................ 4-2 4.2 Summary of Receptor Survey Findings ................................................................. 4-3 Public Water Supply Wells .............................................................................. 4-4 4.2.1 Private Water Supply Wells ............................................................................ 4-4 4.2.2 4.3 Private Water Well Sampling .................................................................................. 4-6 4.4 Numerical Well Capture Zone Analysis ............................................................... 4-8 4.5 Surface Water Receptors .......................................................................................... 4-8 5.0 REGIONAL GEOLOGY AND HYDROGEOLOGY ............................................... 5-1 5.1 Regional Geology ...................................................................................................... 5-1 5.2 Regional Hydrogeology ........................................................................................... 5-3 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page iii P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx TABLE OF CONTENTS SECTION PAGE 6.0 SITE GEOLOGY AND HYDROGEOLOGY ............................................................ 6-1 6.1 Site Geology ............................................................................................................... 6-2 Soil Classification .............................................................................................. 6-2 6.1.1 Rock Lithology .................................................................................................. 6-3 6.1.2 Structural Geology ............................................................................................ 6-5 6.1.3 Soil and Rock Mineralogy and Chemistry .................................................... 6-6 6.1.4 6.2 Site Hydrogeology .................................................................................................... 6-6 Hydrostratigraphic Layer Development ....................................................... 6-7 6.2.1 Hydrostratigraphic Layer Properties ............................................................. 6-7 6.2.2 6.3 Groundwater Flow Direction .................................................................................. 6-8 6.4 Hydraulic Gradient ................................................................................................. 6-10 6.5 Hydraulic Conductivity ......................................................................................... 6-11 6.6 Groundwater Velocity ............................................................................................ 6-11 6.7 Contaminant Velocity ............................................................................................. 6-12 6.8 Slug Test and Aquifer Test Results ...................................................................... 6-13 6.9 Fracture Trace Study Results ................................................................................. 6-14 7.0 SOIL SAMPLING RESULTS ...................................................................................... 7-1 7.1 Background Soil Data ............................................................................................... 7-1 7.2 Facility Soil Data ....................................................................................................... 7-2 8.0 SEDIMENT RESULTS ................................................................................................. 8-1 8.1 Sediment/Surface Soil Associated with AOWs .................................................... 8-1 8.2 Sediment in Major Water Bodies ............................................................................ 8-2 9.0 SURFACE WATER RESULTS .................................................................................... 9-1 9.1 Discussion of Results for Constituents Without Established 2B Standards ..... 9-3 9.2 Comparison of Exceedances of 2B Standards ....................................................... 9-3 9.3 Discussion of Surface Water Results ...................................................................... 9-4 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page iv P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx TABLE OF CONTENTS SECTION PAGE 10.0 GROUNDWATER SAMPLING RESULTS ............................................................ 10-1 10.1 Background Groundwater Concentrations ......................................................... 10-2 Background Dataset Statistical Analysis ..................................................... 10-3 10.1.1 Piper Diagrams (Comparison to Background) ........................................... 10-6 10.1.2 10.2 Downgradient Groundwater Concentrations..................................................... 10-6 Monitoring Wells Beneath Ash Basin .......................................................... 10-6 10.2.1 Monitoring Wells Downgradient of Ash Basin .......................................... 10-8 10.2.2 Monitoring Wells in Separate Flow Regime ............................................... 10-9 10.2.3 Monitoring Wells East of Rail Line (Separate Flow Regime) ................... 10-9 10.2.4 Piper Diagrams (Comparison to Downgradient/ Separate 10.2.5 Flow Regime) ................................................................................................. 10-10 10.3 Site-Specific Exceedances (Groundwater COIs) ............................................... 10-11 Provisional Background Threshold Values (PBTVs) ............................... 10-11 10.3.1 Applicable Standards ................................................................................... 10-11 10.3.2 Additional Requirements ............................................................................. 10-12 10.3.3 Mayo Plant COIs ........................................................................................... 10-13 10.3.4 11.0 HYDROGEOLOGICAL INVESTIGATION .......................................................... 11-1 11.1 Plume Physical and Chemical Characterization ................................................ 11-1 Plume Physical Characterization .................................................................. 11-1 11.1.1 Plume Chemical Characterization ................................................................ 11-4 11.1.2 11.2 Pending Investigation(s) ...................................................................................... 11-17 12.0 RISK ASSESSMENT .................................................................................................. 12-1 12.1 Human Health Screening Summary .................................................................... 12-2 12.2 Ecological Screening Summary ............................................................................. 12-4 12.3 Private Well Receptor Assessment Update ......................................................... 12-4 12.4 Risk Assessment Update Summary ..................................................................... 12-5 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page v P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx TABLE OF CONTENTS SECTION PAGE 13.0 GROUNDWATER MODELING RESULTS........................................................... 13-1 13.1 Summary of Fate and Transport Model Results................................................. 13-2 Flow Model Construction .............................................................................. 13-3 13.1.1 Transport Model Construction ..................................................................... 13-8 13.1.2 Summary of Flow and Transport Modeling Results To Date ................ 13-11 13.1.3 13.2 Summary of Geochemical Model Results ......................................................... 13-13 Model Construction ...................................................................................... 13-13 13.2.1 Summary of Geochemical Model Results To Date................................... 13-16 13.2.2 13.3 Summary of Groundwater to Surface Water Evaluation ................................ 13-18 14.0 SITE ASSESSMENT RESULTS ................................................................................ 14-1 14.1 Nature and Extent of Contamination ................................................................... 14-1 14.2 Maximum COI Concentrations ............................................................................. 14-4 14.3 Contaminant Migration and Potentially Affected Receptors ........................... 14-6 15.0 CONCLUSIONS AND RECOMMENDATIONS ................................................. 15-1 15.1 Overview of Site Conditions at Specific Source Areas ...................................... 15-1 15.2 Revised Site Conceptual Model ............................................................................ 15-1 15.3 Interim Monitoring Program ................................................................................. 15-3 IMP Implementation ....................................................................................... 15-3 15.3.1 IMP Reporting ................................................................................................. 15-4 15.3.2 15.4 Preliminary Evaluation of Corrective Action Alternatives............................... 15-4 CAP Preparation Process ............................................................................... 15-5 15.4.1 Summary .......................................................................................................... 15-7 15.4.2 16.0 REFERENCES ............................................................................................................... 16-1 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page vi P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx LIST OF FIGURES Executive Summary Figure ES-1 Approximate Extent of Impact 1.0 Introduction Figure 1-1 Site Location Map 2.0 Site History and Description Figure 2-1 Mayo Plant Vicinity Map Figure 2-2 1968 USGS Topographic Map Figure 2-3 1948 Aerial Photograph Figure 2-4 1981 Aerial Photograph Figure 2-5 1993 Aerial Photograph Figure 2-6 2008 Aerial Photograph Figure 2-7 NPDES Flow Diagram Figure 2-8 Site Layout Map 3.0 Source Characteristics Figure 3-1 Known Sample of Ash for Comparison Figure 3-2 Elemental Composition for Bottom Ash, Fly Ash, Shale, and Volcanic Ash Figure 3-3 Coal Ash TCLP Leachate Concentration Ranges Compared to Regulatory Limits 4.0 Receptor Information Figure 4-1 USGS Map with Water Supply Wells Figure 4-2 Water Supply Well Locations and Data Figure 4-3 Water Supply Well Capture Zone 5.0 Regional Geology and Hydrogeology Figure 5-1 Regional Geologic Map Figure 5-2 Piedmont Slope-Aquifer System 6.0 Site Geology and Hydrogeology Figure 6-1 Generalized Geologic Map Figure 6-2 General Cross Section A-A' Figure 6-3 General Cross Section B-B' 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page vii P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx LIST OF FIGURES (CONTINUED) Figure 6-4 Generalized Water Level Map - November 2016 Figure 6-5 Surficial Water Level Map - November 2016 Figure 6-6 Transition Zone Water Level Map - November 2016 Figure 6-7 Bedrock Water Level Map - November 2016 Figure 6-8 Generalized Water Level Map - June 2017 Figure 6-9 Surficial Water Level Map - June 2017 Figure 6-10 Transition Zone Water Level Map - June 2017 Figure 6-11 Bedrock Water Level Map - June 2017 Figure 6-12 Potential Vertical Gradient Between Shallow and Deep Zones Figure 6-13 Fracture Trace Analysis 7.0 Soil Sampling Results Figure 7-1 Potential Secondary Source Soil Analytical Results 9.0 Surface Water Results Figure 9-1 Piper Diagram - Surface Water and AOWs 10.0 Groundwater Sampling Results Figure 10-1 Piper Diagram – Surficial Groundwater Figure 10-2 Piper Diagram - Transition Zone Groundwater Figure 10-3 Piper Diagram – Bedrock Groundwater 11.0 Hydrogeological Investigation Figure 11-1 Isoconcentration Map - Arsenic in Surficial Groundwater Figure 11-2 Isoconcentration Map - Arsenic in Transition Zone Groundwater Figure 11-3 Isoconcentration Map - Arsenic in Bedrock Groundwater Figure 11-4 Isoconcentration Map - Barium in Surficial Groundwater Figure 11-5 Isoconcentration Map - Barium in Transition Zone Groundwater Figure 11-6 Isoconcentration Map - Barium in Bedrock Groundwater Figure 11-7 Isoconcentration Map - Boron in Surficial Groundwater Figure 11-8 Isoconcentration Map - Boron in Transition Zone Groundwater Figure 11-9 Isoconcentration Map - Boron in Bedrock Groundwater Figure 11-10 Isoconcentration Map - Calcium in Surficial Groundwater Figure 11-11 Isoconcentration Map - Calcium in Transition Zone Groundwater Figure 11-12 Isoconcentration Map - Calcium in Bedrock Groundwater Figure 11-13 Isoconcentration Map - Chloride in Surficial Groundwater 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page viii P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx LIST OF FIGURES (CONTINUED) Figure 11-14 Isoconcentration Map - Chloride in Transition Zone Groundwater Figure 11-15 Isoconcentration Map - Chloride in Bedrock Groundwater Figure 11-16 Isoconcentration Map - Chromium (VI) and Chromium (Total) in Surficial Groundwater Figure 11-17 Isoconcentration Map - Chromium (Total and Hexavalent) in Transition Zone Groundwater Figure 11-18 Isoconcentration Map - Chromium (Total and Hexavalent) in Bedrock Groundwater Figure 11-19 Isoconcentration Map - Cobalt in Surficial Groundwater Figure 11-20 Isoconcentration Map - Cobalt in Transition Zone Groundwater Figure 11-21 Isoconcentration Map - Cobalt in Bedrock Groundwater Figure 11-22 Isoconcentration Map - Iron in Surficial Groundwater Figure 11-23 Isoconcentration Map - Iron in Transition Zone Groundwater Figure 11-24 Isoconcentration Map - Iron in Bedrock Groundwater Figure 11-25 Isoconcentration Map - Manganese in Surficial Groundwater Figure 11-26 Isoconcentration Map - Manganese in Transition Zone Groundwater Figure 11-27 Isoconcentration Map - Manganese in Bedrock Groundwater Figure 11-28 Isoconcentration Map - Molybdenum in Surficial Groundwater Figure 11-29 Isoconcentration Map - Molybdenum in Transition Zone Groundwater Figure 11-30 Isoconcentration Map - Molybdenum in Bedrock Groundwater Figure 11-31 Isoconcentration Map - Strontium in Surficial Groundwater Figure 11-32 Isoconcentration Map - Strontium in Transition Zone Groundwater Figure 11-33 Isoconcentration Map - Strontium in Bedrock Groundwater Figure 11-34 Isoconcentration Map - Sulfate in Surficial Groundwater Figure 11-35 Isoconcentration Map - Sulfate in Transition Zone Groundwater Figure 11-36 Isoconcentration Map - Sulfate in Bedrock Groundwater Figure 11-37 Isoconcentration Map - Total Dissolved Solids in Surficial Groundwater Figure 11-38 Isoconcentration Map - Total Dissolved Solids in Transition Zone Groundwater Figure 11-39 Isoconcentration Map - Total Dissolved Solids in Bedrock Groundwater Figure 11-40 Isoconcentration Map - Vanadium in Surficial Groundwater Figure 11-41 Isoconcentration Map - Vanadium in Transition Zone Groundwater 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page ix P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx LIST OF FIGURES (CONTINUED) Figure 11-42 Isoconcentration Map - Vanadium in Bedrock Groundwater Figure 11-43 Isoconcentration Map - pH in Surficial Groundwater Figure 11-44 Isoconcentration Map - pH in Transition Zone Groundwater Figure 11-45 Isoconcentration Map - pH in Bedrock Groundwater Figure 11-46 Concentration Versus Distance From Source pH, Boron, Strontium, Sulfate, TDS, Arsenic Constituents Figure 11-47 Concentration Versus Distance From Source Barium, Chromium (VI), Chromium, Cobalt, Iron, Manganese Figure 11-48 Concentration Versus Distance From Source Molybdenum, Vanadium, Radium Figure 11-49 Arsenic Analytical Results - Cross Section C-C' Figure 11-50 Barium Analytical Results - Cross Section C-C' Figure 11-51 Boron Analytical Results - Cross Section C-C' Figure 11-52 Calcium Analytical Results - Cross Section C-C' Figure 11-53 Chloride Analytical Results - Cross Section C-C' Figure 11-54 Chromium (Hexavalent and Total) Analytical Results - Cross Section C-C' Figure 11-55 Cobalt Analytical Results - Cross Section C-C' Figure 11-56 Iron Analytical Results - Cross Section C-C' Figure 11-57 Manganese Analytical Results - Cross Section C-C' Figure 11-58 Molybdenum Analytical Results - Cross Section C-C' Figure 11-59 pH Analytical Results - Cross Section C-C' Figure 11-60 Strontium Analytical Results - Cross Section C-C' Figure 11-61 Sulfate Analytical Results - Cross Section C-C' Figure 11-62 Total Dissolved Solids Analytical Results - Cross Section C-C' Figure 11-63 Vanadium Analytical Results - Cross Section C-C' 12.0 Screening-Level Risk Assessment Figure 12-1 Exposure Areas - Human Health Risk Assessment Figure 12-2 Exposure Areas - Ecological Risk Assessment 14.0 Site Assessment Results Figure 14-1 Time Versus Concentration - pH in Surficial Zone Figure 14-2 Time Versus Concentration - pH in Transition Zone Figure 14-3 Time Versus Concentration - pH in Bedrock Figure 14-4 Time Versus Concentration - Arsenic in Surficial Zone 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page x P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx LIST OF FIGURES (CONTINUED) Figure 14-5 Time Versus Concentration - Arsenic in Transition Zone Figure 14-6 Time Versus Concentration - Arsenic in Bedrock Figure 14-7 Time Versus Concentration - Barium in Surficial Zone Figure 14-8 Time Versus Concentration - Barium in Transition Zone Figure 14-9 Time Versus Concentration - Barium in Bedrock Figure 14-10 Time Versus Concentration - Boron in Surficial Zone Figure 14-11 Time Versus Concentration - Boron in Transition Zone Figure 14-12 Time Versus Concentration - Boron in Bedrock Figure 14-13 Time Versus Concentration - Chromium (Total and Hexavalent) in Surficial Zone Figure 14-14 Time Versus Concentration - Chromium (Total and Hexavalent) in Transition Zone Figure 14-15 Time Versus Concentration - Chromium (Total and Hexavalent) in Bedrock Figure 14-16 Time Versus Concentration - Cobalt in Surficial Zone Figure 14-17 Time Versus Concentration - Cobalt in Transition Zone Figure 14-18 Time Versus Concentration - Cobalt in Bedrock Figure 14-19 Time Versus Concentration - Iron in Surficial Zone Figure 14-20 Time Versus Concentration - Iron in Transition Zone Figure 14-21 Time Versus Concentration - Iron in Bedrock Figure 14-22 Time Versus Concentration - Manganese in Surficial Zone Figure 14-23 Time Versus Concentration - Manganese in Transition Zone Figure 14-24 Time Versus Concentration - Manganese in Bedrock Figure 14-25 Time Versus Concentration - Molybdenum in Surficial Zone Figure 14-26 Time Versus Concentration - Molybdenum in Transition Zone Figure 14-27 Time Versus Concentration - Molybdenum in Bedrock Figure 14-28 Time Versus Concentration - Sulfate in Surficial Zone Figure 14-29 Time Versus Concentration - Sulfate in Transition Zone Figure 14-30 Time Versus Concentration - Sulfate in Bedrock Figure 14-31 Time Versus Concentration - Total Dissolved Solids in Surficial Zone Figure 14-32 Time Versus Concentration - Total Dissolved Solids in Transition Zone Figure 14-33 Time Versus Concentration - Total Dissolved Solids in Bedrock Figure 14-34 Time Versus Concentration - Vanadium in Surficial Zone Figure 14-35 Time Versus Concentration - Vanadium in Transition Zone 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page xi P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx LIST OF FIGURES (CONTINUED) Figure 14-36 Time Versus Concentration - Vanadium in Bedrock Figure 14-37 Time Versus Concentration - Strontium in Surficial Zone Figure 14-38 Time Versus Concentration - Strontium in Transition Zone Figure 14-39 Time Versus Concentration - Strontium in Bedrock Figure 14-40 Groundwater Concentration Trend Analysis – Arsenic in All Flow Layers Figure 14-41 Groundwater Concentration Trend Analysis – Barium in All Flow Layers Figure 14-42 Groundwater Concentration Trend Analysis – Boron in All Flow Layers Figure 14-43 Groundwater Concentration Trend Analysis – Calcium in All Flow Layers Figure 14-44 Groundwater Concentration Trend Analysis – Chloride in All Flow Layers Figure 14-45 Groundwater Concentration Trend Analysis - Chromium (VI and Total) in All Flow Layers Figure 14-46 Groundwater Concentration Trend Analysis – Cobalt in All Flow Layers Figure 14-47 Groundwater Concentration Trend Analysis - Iron in All Flow Layers Figure 14-48 Groundwater Concentration Trend Analysis – Manganese in All Flow Layers Figure 14-49 Groundwater Concentration Trend Analysis – Molybdenum in All Flow Layers Figure 14-50 Groundwater Concentration Trend Analysis – Strontium in All Flow Layers Figure 14-51 Groundwater Concentration Trend Analysis – Sulfate in All Flow Layers Figure 14-52 Groundwater Concentration Trend Analysis - Total Dissolved Solids in All Flow Layers Figure 14-53 Groundwater Concentration Trend Analysis – Vanadium in All Flow Layers Figure 14-54 Groundwater Concentration Trend Analysis – pH in All Flow Layers Figure 14-55 Groundwater Concentration Trend Analysis – Arsenic in Surface Water 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page xii P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx LIST OF FIGURES (CONTINUED) Figure 14-56 Groundwater Concentration Trend Analysis - Barium in Surface Water Figure 14-57 Groundwater Concentration Trend Analysis - Boron in Surface Water Figure 14-58 Groundwater Concentration Trend Analysis – Calcium in Surface Water Figure 14-59 Groundwater Concentration Trend Analysis – Chloride in Surface Water Figure 14-60 Groundwater Concentration Trend Analysis - Chromium (IV and Total) in Surface Water Figure 14-61 Groundwater Concentration Trend Analysis – Cobalt in Surface Water Figure 14-62 Groundwater Concentration Trend Analysis - Iron in Surface Water Figure 14-63 Groundwater Concentration Trend Analysis – Manganese in Surface Water Figure 14-64 Groundwater Concentration Trend Analysis - Molybdenum in Surface Water Figure 14-65 Groundwater Concentration Trend Analysis - Strontium in Surface Water Figure 14-66 Groundwater Concentration Trend Analysis - Sulfate in Surface Water Figure 14-67 Groundwater Concentration Trend Analysis - Total Dissolved Solids in Surface Water Figure 14-68 Groundwater Concentration Trend Analysis - Vanadium in Surface Water Figure 14-69 Groundwater Concentration Trend Analysis – pH in Surface Water Figure 14-70 Comprehensive Groundwater Data Figure 14-71 Comprehensive Surface Water / Area of Wetness Data Figure 14-72 Comprehensive Soil and Sediment Data 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page xiii P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx LIST OF TABLES 2.0 Site History and Description Table 2-1 Well Construction Data Table 2-2 NPDES Groundwater Monitoring Requirements 3.0 Source Characteristics Table 3-1 Ash, Rock, and Soil Composition Table 3-2 Physical Properties of Ash Table 3-3 Mineralogy of Ash Table 3-4 Whole Rock Metal Oxide Analysis of Ash Table 3-5 Whole Rock Elemental Analysis of Ash 6.0 Site Geology Table 6-1 Physical Properties of Soil Table 6-2 Chemical Properties of Soil Table 6-3 Soil, Sediment, and Ash Analytical Methods Table 6-4 Ash Pore Water, Groundwater, Surface Water, and Seep Analytical Methods Table 6-5 Historical Water Level Data Table 6-6 Horizontal Hydraulic Gradient and Flow Velocities Table 6-7 Vertical Hydraulic Gradients Table 6-8 In-situ Hydraulic Conductivity Results Table 6-9 Vertical Hydraulic Conductivity of Undisturbed Soil Samples 7.0 Soil Sampling Results Table 7-1 Provisional Background Threshold Values for Soil Table 7-2 Secondary Soil Evaluation 10.0 Groundwater Sampling Results Table 10-1 Background Groundwater Results Table 10-2 Groundwater Provisional Background Threshold Values 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page xiv P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx LIST OF TABLES (CONTINUED) 13.0 Groundwater Modeling Results Table 13-1 Summary of Kd Values From Batch and Column Studies 15.0 Conclusions and Recommendations Table 15-1 Groundwater Interim Monitoring Program Analytical Methods Table 15-2 Interim Groundwater Monitoring Plan Sample Locations 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page xv P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx LIST OF APPENDICES Appendix A Regulatory Correspondence NCDEQ Expectations Document (July 18. 2017) Completed NCDEQ CSA Update Expectations Check List – Mayo Steam Electric Plant Zimmerman To Draovitch (September 1, 2017) NCDEQ Background Dataset Review (July 7, 2017) Revised Interim Monitoring Plans for 14 Duke Energy Facilities (October 19, 2017) NCDENR NORR Letter (August 13, 2014) 1981 C&D Landfill Appendix B Comprehensive Data Table Comprehensive Data Table Notes Table 1 Groundwater Analytical Results Table 2 Surface Water Results Table 3 AOW Results Table 4 Soil and Ash Results Table 5 Sediment Results Table 6 SPLP Results Appendix C Site Assessment Data CSA Data Reports (Physical) UNCC Soil Sorption Report UNCC Soil Sorption Report Addendum Slug Test Results Appendix D Receptor Surveys Drinking Water Well And Receptor Survey – Mayo Steam Electric Plant (SynTerra, September 2014) Supplement To Drinking Water Well And Receptor Survey – Mayo Steam Electric Plant (SynTerra, November 2014) Update To Drinking Water Well And Receptor Survey – Mayo Steam Electric Plant (SynTerra, September 2016) Dewberry Potable Water Programmatic Evaluation 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page xvi P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx LIST OF APPENDICES Appendix E Supporting Documents STANTEC Draft Report WSP Drawings Appendix F Boring Logs and Well Construction Diagrams Appendix G Methodology Duke Energy Low Flow Sampling Plan (June 10, 2015) Assessment Methodology Appendix H Background Statistical Evaluation Report Appendix I Lab Reports 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page xvii P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx LIST OF ACRONYMS 2B NCDEQ Title 15A, Subchapter 2B. Surface Water and Wetland Standards 2L NCDEQ Title 15A, Subchapter 2L. Groundwater Classification and Standards ADD Average Daily Dose AOW Areas of Wetness ASTM American Society for Testing and Materials C&D Construction and Demolition CAMA Coal Ash Management Act CAP Corrective Action Plan CCR Coal Combustion Residuals CFR Code of Federal Register COI Constituent of Interest COPC Constituents of Potential Concern CSA Comprehensive Site Assessment CSM Conceptual Site Model CP&L Carolina Power & Light DEP Duke Energy Progress, LLC DO Dissolved Oxygen DPT Direct Push Technology DWR Division of Water Resources EDR Environmental Data Resources, Inc. EDXRF Energy Dispersive X-ray Diffraction EMP Effectiveness Monitoring Plan EPC Exposure Point Concentration FGD Flue Gas Desulfurization GAP Groundwater Assessment Work Plan, as amended HAO Hydroxide Phases of Aluminum HMO Hydroxide Phases of Manganese HFO Hydroxide Phases of Iron HI Hazard Index IHSB Inactive Hazardous Site Branch 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page xviii P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx LIST OF ACRONYMS (CONTINUED) IMAC Interim Maximum Allowable Concentrations IMP Interim Monitoring Plan LIDAR Light Detection and Ranging LOAEL lowest-observed-adverse-effect level MCL Maximum Contaminant Level NCAC North Carolina Administrative Code NCDEQ North Carolina Department of Environmental Quality NORR Notice of Regulatory Requirements NCDEQ-DWM North Carolina Department of Environmental Quality – Division of Waste Management NCDEQ-DWR North Carolina Department of Environmental Quality – Division of Water Resources NPDES National Pollutant Discharge Elimination System NTU Nephelometric Turbidity Unit NURE National Uranium Resource Evaluation OSB Oriented Strand Board Plant/Site Mayo Steam Electric Plant PBTV Provisional Background Threshold Value PMCL Primary Maximum Contaminant Level POG Protection of Groundwater PSRG North Carolina Preliminary Soil Remediation Goals PTO Permit to operate PWR Partially Weathered Rock RBC Risk-Based Concentration SCM Site Conceptual Model SMCL Secondary Maximum Contaminant Level SPLP Synthetic Precipitation Leaching Procedure TCLP Toxicity Characteristic Leaching Procedure TDS Total Dissolved Solids TOC Total Organic Carbon TRV Toxicity Reference Values USEPA United States Environmental Protection Agency 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page xix P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx LIST OF ACRONYMS (CONTINUED) UNCC University of North Carolina - Charlotte USGS United States Geological Survey UTL upper tolerance limit 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 1-1 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx 1.0 INTRODUCTION Duke Energy Progress, LLC (Duke Energy, DEP) owns and operates the Mayo Steam Electric Plant (the Mayo Plant, Plant, or Site) located at 10660 Boston Road in Roxboro, Person County, North Carolina (Figure 1-1). The Mayo Plant began operations in 1983 with a single 727 megawatt capacity generating coal-fired unit. Coal combustion residuals (CCR) have historically been managed in the Plant’s on-site ash basin (surface impoundment). CCR were initially deposited in the ash basin by hydraulic sluicing operations. CCR was managed at the Plant’s on-site ash basin and transported via wet sluicing until 2013 when the Mayo Plant converted to a dry ash system in which 90 percent of CCR was dry. Final system upgrades were completed in October 2016; all CCR collection is dry. Beginning in November 2014, CCR from the Plant has been managed in a newly constructed on-site Industrial landfill (monofill) permitted by the North Carolina Department of Environment Quality (NCDEQ)1 Division of Waste Management (NCDEQ-DWM) Permit 7305. Discharge from the ash basin to Mayo Lake is permitted by the NCDEQ Division of Water Resources (NCDEQ-DWR) under National Pollutant Discharge Elimination System (NPDES) Permit NC0038377. 1.1 Purpose of Comprehensive Site Assessment This Comprehensive Site Assessment (CSA) update was conducted to refine and expand the understanding of subsurface geologic/hydrogeologic conditions and evaluate the extent of impacts from historical management of coal ash in the ash basin. This CSA update contains an assessment of Site conditions based on a comprehensive interpretation of geologic and sampling results from the initial Site assessment and geologic and sampling results obtained subsequent to the initial assessment and has been prepared in coordination with Duke Energy and NCDEQ in response to requests for additional information, including additional sampling and assessment of specified areas. This CSA update was prepared in conformance to the most recently updated CSA table of contents provided by NCDEQ to Duke Energy on September 29, 2017. In response to a request from NCDEQ for an updated CSA report, this submittal includes the following information. The NCDEQ Expectations Document (July 18, 2017) and the completed NCDEQ CSA Update Expectations Check List are included in Appendix A: 1 Prior to September 18, 2015, the NCDEQ was referred to as the North Carolina Department of Environment and Natural Resources (NCDENR). Both naming conventions are used in this report, as appropriate. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 1-2 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx Review of baseline assessment data collected and reported as part of CSA activities; A summary of NPDES and Coal Ash Management Act (CAMA) groundwater monitoring information; A summary of potential receptors including results from water supply wells; A description and findings of additional assessment activities conducted since submittal of the CSA Supplement report(s); An update on background concentrations for groundwater and soil; Definition of horizontal and vertical extent of CCR constituents in soil and groundwater based upon NCDEQ approved background concentrations; and An update to human health and ecological risk assessment to evaluate the existence of imminent hazards to public health, safety and the environment. 1.2 Regulatory Background The CAMA of 2014 directs owners of CCR surface impoundments in North Carolina to conduct groundwater monitoring, assessment, and remedial activities, if necessary. The CSA was performed to collect information necessary to evaluate the horizontal and vertical extent of impacts to soil and groundwater attributable to CCR source area(s), identify potential receptors, and screen for potential risks to those receptors. Notice of Regulatory Requirements (NORR) 1.2.1 On August 13, 2014, North Carolina Department of Environment and Natural Resources (NCDENR) issued a Notice of Regulatory Requirements (NORR) letter notifying Duke Energy that exceedances of groundwater quality standards were reported at 14 coal ash facilities owned and operated by Duke Energy. Those groundwater quality standards are part of 15A NCAC 02L (2L) .0200 Classifications and Water Quality Standards Applicable to the Groundwaters of North Carolina. The NORR stipulated that for each coal ash facility, Duke Energy was to conduct a CSA. The NORR also stipulated that before conducting each CSA, Duke was to submit a Groundwater Assessment Work Plan (GAP or Work Plan) and a receptor survey. In accordance with the NORR requirements, a receptor survey was performed to identify all receptors within a 0.5-mile radius (2,640 feet) of the ash basin compliance boundary, and a CSA was conducted for each facility. The NORR letter is included in Appendix A. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 1-3 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx Coal Ash Management Act Requirements 1.2.2 The CAMA of 2014 — General Assembly of North Carolina Senate Bill 729 Ratified Bill (Session 2013) (SB 729) requires that ash from Duke Energy coal plant sites located in North Carolina either (1) be excavated and relocated to fully lined storage facilities or (2) go through a classification process to determine closure options and schedule. Closure options can include a combination of excavating and relocating ash to a fully lined structural fill, excavating and relocating the ash to a lined landfill (on-site or off-site), and/or capping the ash with an engineered synthetic barrier system, either in place or after being consolidated to a smaller area on-site. As a component of implementing this objective, CAMA provides instructions for owners of coal combustion residuals surface impoundments to perform various groundwater monitoring and assessment activities. Section §130A-309.209 of the CAMA ruling specifies groundwater assessment and corrective actions, drinking water supply well surveys and provisions of alternate water supply, and reporting requirements as follows: (a) Groundwater Assessment of Coal Combustion Residuals Surface Impoundments. – The owner of a coal combustion residuals surface impoundment shall conduct groundwater monitoring and assessment as provided in this subsection. The requirements for groundwater monitoring and assessment set out in this subsection are in addition to any other groundwater monitoring and assessment requirements applicable to the owners of coal combustion residuals surface impoundments. (1) No later than December 31, 2014, the owner of a coal combustion residuals surface impoundment shall submit a proposed Groundwater Assessment Plan for the impoundment to the Department for its review and approval. The Groundwater Assessment Plan shall, at a minimum, provide for all of the following: a. A description of all receptors and significant exposure pathways. b. An assessment of the horizontal and vertical extent of soil and groundwater contamination for all contaminants confirmed to be present in groundwater in exceedance of groundwater quality standards. c. A description of all significant factors affecting movement and transport of contaminants. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 1-4 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx d. A description of the geological and hydrogeological features influencing the chemical and physical character of the contaminants. 2) The Department shall approve the Groundwater Assessment Plan if it determines that the Plan complies with the requirements of this subsection and will be sufficient to protect public health, safety, and welfare; the environment; and natural resources. (3) No later than 10 days from approval of the Groundwater Assessment Plan, the owner shall begin implementation of the Plan. (4) No later than 180 days from approval of the Groundwater Assessment Plan, the owner shall submit a Groundwater Assessment Report to the Department. The Report shall describe all exceedances of groundwater quality standards associated with the impoundment. 1.3 Approach to Comprehensive Site Assessment This CSA has been performed to meet NCDEQ requirements associated with potential site remedy selection. The following components were utilized to develop the assessment. NORR Guidance 1.3.1 The NORR requires that site assessment provide information to meet the requirements of 2L .0106 (g). This regulation lists the items to be included in site assessments conducted pursuant to Paragraph (c) of the rule. These requirements are listed below and referenced to the applicable sections of this CSA. 15A NCAC 02L .0106(g) Requirement CSA Section(s) (1) The source and cause of contamination Section 3.0 (2) Any imminent hazards to public health and safety, as defined in G.S. 130A-2, and any actions taken to mitigate them in accordance with Paragraph (f) of this Rule Sections ES.2 and 2.8 (3) All receptors and significant exposure pathways Sections 4.0 and 12.0 (4) The horizontal and vertical extent of soil and groundwater contamination and all significant factors affecting contaminant transport Sections 7.0, 8.0, and 14.0 (5) Geological and hydrogeological features influencing the movement, chemical, and physical character of the contaminants Sections 6.0, 11.0, and 15.0 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 1-5 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx USEPA Monitored Natural Attenuation Tiered Approach 1.3.2 The assessment data is compiled in a manner to be consistent with “Monitored Natural Attenuation of Inorganic Contaminants in Groundwater” (EPA/600/R- 07/139) (USEPA, October 2007). The tiered analysis approach discussed in this guidance document is designed to align site characterization tasks to reduce uncertainty in remedy selection. The tiered assessment data collection includes information to evaluate: Active contaminant removal from groundwater and dissolved plume stability, The mechanisms and rates of attenuation, The long-term capacity for attenuation and stability of immobilized contaminants, and Anticipated performance monitoring needs to support the selected remedy. ASTM Conceptual Site Model Guidance 1.3.3 The American Society for Testing and Materials (ASTM) E1689-95 generally describes the major components of conceptual site models, including an outline for developing models. To the extent possible, this guidance was incorporated into preparation of the Site Conceptual Model (SCM). The SCM is used to integrate Site information, identify data gaps, and determine whether additional information is needed at the Site. The model is also used to facilitate selection of remedial alternatives and effectiveness of remedial actions in reducing the exposure of environmental receptors to contaminants (ASTM, 2014). 1.4 Technical Objectives The rationale for CSA activities fall into one of the following categories: Determine the range of background groundwater quality from pertinent geologic settings (horizontal and vertical) across a broad area of the Site. Evaluate groundwater quality from pertinent geologic settings (horizontal and vertical extent of CCR leachate constituents). Establish perimeter (horizontal and vertical) boundary conditions for groundwater modeling. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 1-6 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx Provide source area information including ash pore water chemistry, physical and hydraulic properties, CCR thickness, and residual saturation within the ash basin. Address soil chemistry in the vicinity of the ash basin (horizontal and vertical extent of CCR leachate constituents in soil) compared to background concentrations. Determine potential routes of exposure and receptors. Compile information necessary to develop a groundwater Corrective Action Plan (CAP) protective of human health and the environment in accordance with 2L. 1.5 Previous Submittals Detailed descriptions of the Site operational history, the Site conceptual model, physical setting and features, geology/hydrogeology, and results of the findings of the CSA and other CAMA-related work are documented in full in the following: Comprehensive Site Assessment Report - Mayo Steam Electric Plant (SynTerra, 2015a). Corrective Action Plan Part 1 - Mayo Steam Electric Plant (SynTerra, 2015b). Corrective Action Plan Part 2 – Mayo Steam Electric Plant (SynTerra, 2016a). Comprehensive Site Assessment Supplement 1 – Mayo Steam Electric Plant (SynTerra, 2016b). 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 2-1 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx 2.0 SITE HISTORY AND DESCRIPTION An overview of the Mayo Steam Electric Plant setting and operations is presented in the following sections. 2.1 Site Description, Ownership and Use History The Mayo Plant is a coal-fired electricity-generating facility in north-central North Carolina. The Plant is in the northeastern corner of Person County, north of the City of Roxboro. The address of the Mayo Plant is 10660 Boston Road (US Highway 501), Roxboro, North Carolina. The Carolina Power & Light (CP&L) Mayo Plant became fully operational in June 1983. CP&L merged with Florida Progress Corporation in 2000 to become Progress Energy Inc. Progress Energy merged with Duke Energy in July 2012. The Mayo Plant began coal-fired power production in 1983. A scrubber system is currently in place to reduce emissions from coal combustion. At the Site, there is a single ash basin located northwest of the Plant. The ash basin, which contains ash generated from the Plant’s historic coal combustion, is approximately 140 acres in size, is constructed with an earthen dike. A 500-foot compliance boundary for the routine groundwater monitoring well network encircles the ash basin, except on the northeastern edge of the Site where the compliance boundary is co-located with the boundary of the Site. No other areas of coal ash, other than possible de minimis quantities, are known to exist at the Site other than in the lined landfill described below. CCR was managed at the Plant’s on-site ash basin and transported via wet sluicing until 2013. That year, the Mayo Plant converted to a dry ash system in which 90 percent of CCR was dry. After the conversion, wet sluicing was conducted when there was a shutdown of the dry fly ash collection system. Final system upgrades were completed in October 2016; all CCR collection is dry. Until recently, dry ash had been hauled and disposed in the lined landfill located at the nearby Roxboro Steam Electric Plant (near Semora, North Carolina). Beginning in November 2014, CCR from the Mayo Plant has been placed in a newly constructed on-site industrial landfill (monofill) (NCDEQ DWM Permit No. 7305) at the Mayo Site. A Flue Gas Desulfurization (FGD) system is active at the Mayo Plant. The FGD system directs flue gas into an absorber where limestone (calcium carbonate) slurry is sprayed. Sulfur dioxide in the flue gas reacts with the limestone slurry to produce calcium sulfate, or gypsum. The system reclaims the unreacted limestone slurry to be reused in the absorber. A small blowdown stream is used to maintain the chloride concentration in the reaction tank. There are two FGD ponds at the Mayo Site. The FGD Forward 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 2-2 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx Flush Pond was originally used in the bioreactor treatment process. The bioreactor has been decommissioned and the FGD Forward Flush Pond is inactive; it no longer receives the back-flush of the bioreactor. The FGD Settling Pond receives the stream of FGD blowdown water as well as leachate from the monofill. From the FGD settling pond the water is pumped to the thermal evaporator system. The thermal evaporator system is a process that creates a clean distillate and brine. The brine solution is mixed with fly ash that is placed in the on-site monofill. The clean distillate is used in absorber make-up water. The FGD ponds were constructed within the footprint of the ash basin; however, both ponds are constructed with an engineered liner system. 2.2 Geographic Setting, Surrounding Land Use, Surface Water Classification The Mayo Plant is situated in a rural area in the northeast corner of Person County, approximately 10 miles north of Roxboro, North Carolina and approximately 13 miles south of South Boston, Virginia on US Highway 501 (Boston Road; Figure 1-1). A description of the physical setting of the Mayo Site is described in the following sections. Geographic Setting The primary operational portion of the Site is situated east of US Highway 501 and includes the power plant, the majority of Plant operational facilities and equipment, the ash basin, and Mayo Lake (Figure 2-1). The Site’s northern property line extends to the North Carolina/Virginia state line. The overall topography of the Site generally slopes toward the east (Mayo Lake aka Mayo Reservoir) and northeast (Crutchfield Branch). The Site is roughly bisected by US Highway 501. The majority of the Site - including the power plant, the ash basin, and most of its operational features - is located east of US Highway 501 and extends to the eastern shore of Mayo Lake. The property is bounded to the north by the North Carolina/Virginia state line. Mayo Lake Road runs west to east from the intersection with US Highway 501 to NC county road 1504. The portion of the Site located west of US Highway 501 contains the recently operational monofill and a haul road that connects the monofill with the operational portion of the Mayo Plant. A railroad spur bisects the western portion of the property. RT Hester Road also cuts across the northeastern portion of the western property. The eastern portion of the Site, excluding Mayo Lake, encompasses 460 acres. Mayo Lake encompasses 2,880 acres with a normal water elevation of 432.3 feet 2. 2The datum for all elevation information presented in this report is NAVD88. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 2-3 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx The terrestrial portion of the Site (east) is roughly bisected by a railroad line that supplies the Plant. The ash basin is situated to the northwest of the railroad line, and the power plant and majority of supporting operational features (e.g., coal pile, cooling towers, administrative buildings, substation, etc.) are situated southeast of the railroad line. Forested areas encircle the southeast portion of the property to the southwest, south, east, and northwest. Mayo Lake borders the entire eastern portion of this part of the Site. The ash basin is the dominant feature on the portion of the property northwest of the railroad line and east of US Highway 501. Heavily wooded land surrounds the ash basin with the exception of the wastewater treatment facility on the southeast side of the ash basin. A construction and demolition (C&D) landfill is located east of the ash basin. The permitted landfill (NCDEQ DWM Permit No. 73-B) was used to contain debris generated during the construction of the Plant in the early 1980s. The landfill permit was reopened briefly in the 1990s for a new project; however, the project was cancelled (Appendix A). No additional material was placed in the landfill, and the permit was closed again. The landfill is covered with thick vegetative undergrowth and trees. A large area west of the former C&D landfill, east of the railroad, and south of Mayo Lake Road, has been cleared for construction of a new FGD retention basin. Similarly, a large area on the east side of the Plant haul road, north of the rail line, and just east of the ash basin, has been cleared for the construction of a lined retention pond. Construction activities are ongoing for both locations. Surrounding Land Use Properties located within a 500-foot radius of the Mayo Plant ash basin compliance boundary are all contained within the Site, with the exception of an undeveloped parcel located due north of the northern Plant boundary along Mayo Lake Road. Properties adjacent to the Site are located in Person County, North Carolina and Halifax County, Virginia. Undeveloped land occurs on the Site west of US Highway 501, with the exception of the Mayo monofill. The closest residences to the east of the Site are along the easternmost shore of Mayo Lake. Undeveloped land borders the Site to the north. Several residences are located just outside the Site boundaries to the south and northwest (North Carolina and Virginia). Louisiana Pacific Corporation, located south and west of the Mayo property boundary, at 10475 Boston Road (Roxboro, North Carolina) opposite the Mayo Plant’s entrance road, is a manufacturing facility that produces oriented strand board (OSB). The facility has been in operation since 1994. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 2-4 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx Review of a 1968 topographic map (Figure 2-2) and aerial photograph from 1948 (Figure 2-3) shows the property occupied by forested areas, large cultivated agricultural fields, and a few farms/homes. An aerial photograph from 1981 (Figure 2-4) shows the Plant under construction and Mayo Lake filling with water. The Mayo Plant ash basin dam had not yet been built. In a 1993 aerial photograph (Figure 2-5), the Plant is operational and appears much as it does now. At that time, ash was being actively sluiced into the ash basin, which was largely open water. The 1993 photograph shows that the majority of sluiced ash was discharged to the basin on the northwest side of the railroad line. An obvious “ash delta” with active sluicing is noted in the photographs. A 2008 photograph (Figure 2-6) indicates that the Site appears almost identical to present day, with mostly wooded areas surrounding the ash basin and power plant areas. In that photograph, the ash basin is approximately half open water and half ash. Meteorological Data The Site lies within the Piedmont region of the southeastern United States and exhibits a humid, subtropical climate type (NOAA, 2013). More specifically, the Site lies in the northern Piedmont of North Carolina where the mean annual temperature is about 58 degrees Fahrenheit (F) and average annual precipitation is approximately 44 inches (State Climate Office, 2017). A weather data station maintained by the state is located just east of the city of Roxboro and about 11 miles south of the Mayo Plant. The mean annual temperature recorded in Roxboro is 56.9 degrees F with a minimum annual temperature average of 25.7 degrees F in January and a maximum annual temperature average of 87.8 degrees F in July. Precipitation in Roxboro averages 46.6 inches annually, slightly higher than the average for the northern Piedmont (State Climate Office, 2017). Surface Water Classification The Site is located within the Roanoke River Basin. Major surface water bodies in the vicinity of the Mayo Plant are described in this section. There is no groundwater influence from the ash basin on the surface water bodies described with the possible exception of Crutchfield Branch (Section 9.0). Mayo Lake is the dominant feature on the eastern portion of the Site. Mayo Lake was formed when Mayo Creek was dammed, and it now encapsulates the majority of the reach of Mayo Creek as well as a number of smaller streams that flowed into Mayo Creek prior to the lake formation. The 2,800-acre lake is maintained by the earthen dam located at the northern end of the lake. The dam is approximately 100 feet high. Water level is controlled by a spillway located on the eastern end of the lake/dam. A diversion/drawdown pipe (72-inch diameter) also discharges water at the base of the 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 2-5 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx dam on the bottom of the eastern end of the dam. The water flowing from this pipe merges with the water from the spillway to reform Mayo Creek. Mayo Creek flows north and then northwest and merges with Crutchfield Branch flowing northeast. The merged Mayo Creek flows north and merges with Hyco River just south of Cluster Springs, Virginia. The Hyco River flows northeast and merges with the Dan River just upstream of Kerr Reservoir (Roanoke River). Mayo Lake and the portion of Mayo Creek that flows into Mayo Lake are classified as WS-V waters. Mayo Creek, as it flows from the Mayo Lake dam to the Virginia state line is classified as “C”. Crutchfield Branch, from its source to the Virginia state line, is classified as “B”. One other “blue line” stream is present on the Site and is designated as Bowes Branch. The stream is classified as “C” waters and flows through the western portion of the Site north toward the state line (NCDENR, 2015). Several intermittent streams are also present on the Site. A channel associated with NPDES Outfall 002 originates on the east side of the ash basin and flows southeast, under the railroad line, and into Mayo Lake. 2.3 CAMA-related Source Areas CAMA provides for groundwater assessment of CCR surface impoundments defined as topographic depressions, excavations, or diked areas formed primarily of earthen materials, without a base liner, and that meet other criteria related to design, usage, and ownership (Section 130A-309.201). At Mayo, the groundwater assessment was conducted for the ash basin CCR surface impoundment. The Mayo Plant ash basin is located in the northern portion of the Site, 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 dam height of 110 feet and a crest height elevation of 479.8 feet. The entire basin area is approximately 140 acres and contains approximately 6,600,000 tons of CCR material (Duke Energy, 2017). Borings installed in the ash basin encountered ash from 13.5 feet to 66.1 feet in thickness. Roughly 40 percent of the ash basin is currently covered with standing water. No other ash storage facilities have been identified on the Site other than the permitted monofill. The Mayo Plant ash basin was constructed by damming Crutchfield Branch. Examination of historic United States Geological Survey (USGS) aerial photography and topographic maps (Figure 2-2 to Figure 2-6) indicates that the Site was heavily wooded with some open pasture prior to use for ash settling. Three tributaries that formed Crutchfield Branch are present within heavily wooded areas in a 1948 aerial photograph (Figure 2-3). A 1981 historical aerial photograph (Figure 2-4) shows the Plant under 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 2-6 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx construction and the dam at Mayo Lake under construction and water being impounded. However, the ash basin dam has not yet been constructed nor has Site clearing for the ash basin been initiated. Construction of the ash basin was complete by October 1982, and the Mayo Plant began operations in June 1983. The ash basin dam is an earthen embankment armored with rip rap on the basin side and on the downslope base of the dam. The perimeter of the basin is mostly unaltered and well-vegetated with the exception of the ash basin dam and a small shoreline section on the east (near the forebay) that are armored with rip rap. The ash basin dam and dam access road are raised about 10 feet higher than the ash basin water level. The Mayo Plant ash basin was constructed with two engineered toe drains located at the base of the dam. On the southeast side of the ash basin, water enters a holding lagoon or forebay for water quality treatment and flows 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. 15A NCAC 02L .0106 (f)(4) requires secondary sources that could be potential continuing sources of possible impact to groundwater be addressed in the CAP. At the Mayo Site, the soil located below the ash basin could be considered a potential CAMA- related secondary source. Information to date indicates that the thickness of soil impacted by ash would generally be limited to the depth interval near the ash/soil interface. 2.4 Other Primary and Secondary Sources CSA activities included an assessment of the horizontal and vertical extent of constituents related to ash management areas observed at concentrations greater than 2L, Interim Maximum Allowable Concentrations (IMAC) or PBTVs. If the CSA indicates constituent exceedances are related to sources other than the ash basin, those sources will be addressed as part of a separate process in compliance with the requirements of 2L. 2.5 Summary of Permitted Activities The NPDES program regulates wastewater discharges to surface waters. The Mayo Plant is permitted to discharge wastewater under NPDES Permit NC0038377, which authorizes discharge from the facility to Mayo Lake in accordance with effluent limitations, monitoring requirements, and other conditions set forth in the permit. Ash basin is referred to as “ash pond” in the Plant’s NPDES permit. The most recent NPDES permit for the Mayo Plant became effective on November 1, 2009, and expired in March 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 2-7 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx 2012. NPDES flow diagrams for the Mayo Plant are included as Figure 2-7. A permit renewal request has been submitted to NCDENR and is pending so the Site continues to operate under the administratively extended permit. NPDES permit NC0038377 authorizes two discharges to Mayo Lake. Outfall 001 discharges cooling tower water and circulating water system discharge water. Outfall 002 is comprised of a number of streams including internal Outfall 008 (cooling tower blowdown); internal Outfall 009 (FGD blowdown); ash transport water; coal pile runoff; and other sources including water from wastewater treatment processes. Nine stormwater outfalls are also authorized for the Mayo Plant. Surface water monitoring from Mayo Lake has been conducted at the Site since the Plant began operations. In its current configuration, the monitoring plan includes sampling for water quality, water chemistry, phytoplankton, chlorophyll, and zebra mussels on an alternating calendar month frequency. Fisheries sampling is conducted four times each year, trace element sampling is conducted once per year, and an aquatic vegetation survey is conducted once per year. In addition to surface water monitoring, the NPDES permit requires groundwater monitoring. Permit Condition A (6) Attachment X, Version 1.0, dated March 17, 2011, lists the groundwater monitoring wells to be sampled, the parameters and constituents to be measured and analyzed, and the requirements for sampling frequency and results reporting. Details are provided in Section 2.6. A C&D landfill is present on the Site. That landfill is located to the east of the ash basin (Figure 2-4). The permitted landfill (NCDEQ DWM; Permit No. 73-B) was used to contain debris generated during the construction of the Mayo Plant in the early 1980s. The landfill permit was reopened briefly in the 1990s for a new project; however, the project was cancelled, no additional placement of material occurred in the landfill, and the permit was closed again. As previously mentioned in the report, beginning in November 2014, CCR from the Mayo Plant has been placed in a newly constructed on-site monofill (NCDEQ DWM; Permit No. 7305). NCDEQ DWM issued a permit to operate (PTO) the monofill on July 10, 2014. The monofill is located on an upland area west of the main Plant area (Figure 2-1). Phase 1 consists of 31 acres out of a possible total 104-acre proposed landfill footprint. The capacity of Phase 1 is 1,592,000 cubic yards. The monofill was constructed with a double high-density polyethylene liner with leak detection, groundwater monitoring, and leachate collection systems. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 2-8 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx 2.6 History of Site Groundwater Monitoring The location of the ash basin voluntary and compliance monitoring wells, CSA wells, the approximate ash basin waste boundary, and the compliance boundary are shown in Figure 2-8. Construction details for Site monitoring wells are provided in Table 2-1. At Mayo, CAMA monitoring wells are designated with an S, D, BR, or BRL identifier. These designations correspond to the flow unit in which the well is screened. “S” refers to the surficial flow layer (alluvium and saprolite); “D” refers to the poorly weathered rock within a transition zone between the surficial flow zone and bedrock; and “BR” and “BRL” refer to the bedrock flow zone (relatively unfractured rock). The following sections discuss groundwater monitoring activities associated with CCR conducted prior to CSA activities through current CAMA-related monitoring activities. Groundwater monitoring results are presented in Section 10.0. Ash Basin Voluntary Groundwater Monitoring 2.6.1 Four monitoring wells were voluntarily installed in 2008 by Duke Energy for the purpose of groundwater monitoring. Voluntary well MW-1 has subsequently been renamed “BG-1” and is used as a background well for the scheduled NPDES groundwater compliance monitoring. Voluntary wells MW-2 and MW-3 are located downgradient of the ash basin dam. Voluntary well MW-4 is located in a sidegradient/upgradient position just east of the ash basin. Wells MW-2, MW-3, and MW-4 are not included in NPDES-related monitoring but have been included in groundwater assessment activities associated with the CSA. Ash Basin NPDES Groundwater Monitoring 2.6.2 The NPDES permit requires routine groundwater monitoring. Permit Condition A (6) Attachment XX, Version 1.0, dated March 17, 2011, lists the groundwater monitoring wells to be sampled, the parameters and constituents to be measured and analyzed, and the requirements for sampling frequency and results reporting. Permit Condition A (6) Attachment XX also provides requirements for well location and well construction. Groundwater monitoring events related to the NPDES permit for the Site are conducted three times per year. The ash basin has a network of 10 monitoring wells which includes (BG-1 (background), BG-2 (background), CW-1, CW-1D, CW-2, CW-2D, CW-3, CW-4, CW-5, and CW-6). Duke Energy initiated routine compliance boundary monitoring for the ash basin in December 2010. Currently, Duke Energy conducts routine NPDES groundwater compliance boundary monitoring during April, July, and November each year for the parameters listed in Table 2-2. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 2-9 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx Ash Basin CAMA Groundwater Monitoring 2.6.3 Thirty-four (34) groundwater wells were installed as part of this assessment (Figure 2-8). Groundwater monitoring wells (MW-12S, MW-12D, and MW-14BR) were installed and are considered background wells. Fourteen groundwater monitoring wells (MW-3BR, MW-5BR, MW-7D, MW-7BR, MW-8BR, MW-9BRL, MW-13BR, MW-16S, MW-16D, MW-16BR, MW-18D, MW-18BR, MW-19D, and MW-19BR) and eight piezometers (MW-6BR, MW-8S, MW-8D, MW-9BR, MW- 10BR, MW-11BR, MW-15BR, MW-17BR) were installed at locations outside of the perimeter of the ash basin. Nine groundwater monitoring wells (ABMW-1, ABMW-2, ABMW-2BR, ABMW-2BRL, ABMW-3, ABMW-3S, ABMW-4, ABMW- 4D, and ABMW-4BR) were installed at four locations in the ash basin. Compliance monitoring wells BG-1, BG-2, CW-1, CW-1D, CW-2, CW-2D, CW-3, CW-4, CW-5, and CW-6, as well as previously installed wells MW-2, MW-3, and MW-4 were also used for this assessment. Landfill Groundwater Monitoring 2.6.4 Duke Energy conducts routine solid waste landfill compliance monitoring in accordance with NCDEQ DWM Permit No. 7305; however, that monitoring is not subject to CAMA, and is not detailed herein. The monofill, as previously described, is located west of the US Highway 501 in a separate groundwater flow regime, and is not influenced by the ash basin. 2.7 Summary of Assessment Activities With the exception of the voluntary and compliance well installation/groundwater monitoring described above, no other known CCR-related groundwater investigations or site assessments were conducted at Mayo Plant prior to implementation of the GAP (SynTerra, 2014c). 2.8 Summary of Initial Abatement, Source Removal or Other Corrective Action No abatement or source removal activities have been conducted at the Mayo Plant related to the ash basin other than discontinuing the discharge of ash and sluice water. As previously indicated, CCR was historically managed at the Plant’s on-site ash basin and transported via wet sluicing. In 2013, the Mayo Plant converted to a dry ash collection system. Additional upgrades to the dry fly ash handling system were completed in October 2016 and all CCR collection is now dry. In preparation for ash basin closure, new retention basins and wastewater treatment systems are being designed and constructed. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 3-1 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx 3.0 SOURCE CHARACTERISTICS For purposes of this assessment, the source area is defined by the ash waste boundary as depicted on Figure 2-8. For the Mayo Site, source areas include the ash management areas comprised of the Ash Basin and two FGD ponds. 3.1 Coal Combustion and Ash Handling System Coal ash is produced from the combustion of coal. The coal is dried, pulverized, and conveyed to the burner area of a boiler. The smaller particles produced by coal combustion, referred to as fly ash, are carried upward in the flue gas and are captured by an air pollution control device, such as an electrostatic precipitator. The larger particles of ash that fall to the bottom of the boiler are referred to as bottom ash or boiler slag. After Plant operations began in 1983, CCR was historically managed at the Plant’s on- site ash basin and transported via wet sluicing. In 2013, the Mayo Plant converted to a dry ash system in which 90 percent of CCR was dry. Final system upgrades were completed in October 2016 so that all CCR collection is dry. Prior to 2014, dry ash had been hauled and disposed in the industrial landfill located at the nearby Roxboro Steam Electric Plant (near Semora, North Carolina). Beginning in November 2014, CCR from the Mayo Plant has been placed in the on-site Mayo monofill. The Mayo Plant began coal-fired power production in 1983. At the Site, there is a single ash basin located northwest of the Plant that contains ash generated from the Plant’s historic coal combustion. The ash basin is approximately 140 acres in size, is constructed with an earthen dike, and contains approximately 6,600,000 tons of CCR (Duke Energy, 2017). The ash basin dam is an earthen embankment armored with rip rap on the basin side and on the downslope base of the dam. The perimeter of the basin is mostly unaltered and well-vegetated with the exception of the ash basin dam and a small shoreline section on the east (near the forebay) that are armored with rip rap. The ash basin dam and dam access road are raised about 10 feet higher than the ash basin water level. Roughly one-half of the ash basin is covered with standing water. A FGD system is active at the Mayo Plant. The FGD system directs flue gas into an absorber where limestone (calcium carbonate) slurry is sprayed. Sulfur dioxide in the flue gas reacts with the limestone slurry to produce calcium sulfate, or gypsum. The system reclaims the unreacted limestone slurry to be reused in the absorber. A small blowdown stream is used to maintain the chloride concentration in the reaction tank. There are two lined FGD ponds at the Mayo Site. The liner system is comprised of - 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 3-2 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx from bottom to top – fill, a geocomposite liner, and a 60-mil HDPE liner. The FGD Forward Flush Pond was originally used in the bioreactor treatment process. The bioreactor has been decommissioned and the FGD Forward Flush Pond is inactive; it no longer receives the back-flush of the bioreactor. The FGD Settling Pond receives the stream of FGD blowdown water as well as leachate water from the monofill. From the FGD settling pond the water is pumped to the thermal evaporator system. The thermal evaporation system is a process that creates a clean distillate and brine. The brine solution is mixed with fly ash that is placed in the on-site monofill. The clean distillate is used in absorber make-up water. 3.2 General Physical and Chemical Properties of Ash Coal ash consists of fly ash and bottom ash produced from the combustion of coal. The physical and chemical properties of coal ash are determined by reactions that occur during the combustion of the coal and subsequent cooling of the flue gas. Physical Properties Approximately 70 percent to 80 percent of the ash produced during coal combustion is fly ash (EPRI, 1993). Typically 65 percent to 90 percent of fly ash has particle sizes that are less than 0.010 millimeter (mm). In general, fly ash has a grain size distribution similar to that of silt. The remaining 20 percent to 30 percent of ash produced is considered bottom ash. Bottom ash consists of angular particles with a porous surface and is normally gray to black in color. Bottom ash particle diameters can vary from approximately 38 mm to 0.05 mm. In general, bottom ash has a grain size distribution similar to that of fine gravel to medium sand (EPRI, 1995). Physical properties of ash are on Table 3-1. Based on published literature not specific to Mayo, the specific gravity of fly ash ranges from 2.1 to 2.9, and the specific gravity of bottom ash typically ranges from 2.3 to 3.0. The permeability of fly ash and bottom ash vary based on material density, but would be within the range of a soil with a similar gradation and density (EPRI, 1995). Chemical Properties The specific mineralogy of coal ash varies based on many factors including the chemical composition of the coal, which is directly related to the geographic region where the coal was mined; the type of boiler where the combustion occurs (i.e., thermodynamics of the boiler); and air pollution control technologies employed. The overall chemical composition of coal ash resembles that of siliceous rocks from which it was derived, particularly shale. Oxides of silicon, aluminum, iron, and calcium 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 3-3 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx make up more than 90 percent of most siliceous rocks, soils, fly ash, and bottom ash. Other major and minor elements (sulfur, sodium, potassium, magnesium, titanium) make up an additional 8 percent, while trace constituents account for less than one percent. The following constituents are considered to be trace elements: arsenic, barium, cadmium, chromium, lead, mercury, selenium, copper, manganese, nickel, lead, vanadium, and zinc (EPRI, 2010). The majority of fly ash particles are glassy spheres mainly composed of amorphous or glassy aluminosilicates, crystalline matter, and carbon. Figure 3-1 presents a photograph of ash collected from the ash basin at Duke Energy's Cliffside Steam Station (considered representative of the ash at the Mayo Plant) showing a mix of fly ash and bottom ash at 10 µm and 20 µm magnifications. The glassy spheres can be observed in the photograph. The glassy spheres are generally immune to dissolution. During the later stages of the combustion process and as the combustion gases are cooling after exiting the boiler, molecules from the combustion process condense on the surface of the glassy spheres. These surface condensates consist of soluble salts [e.g., calcium (Ca2+) and sulfate (SO2-)], metals [e.g., copper (Cu) and zinc (Zn)], and other minor elements [e.g., boron (B), selenium (Se), and arsenic (As)] (EPRI, 1994). The major elemental composition of fly ash (approximately 95 percent by weight) is composed of mineral oxides of silicon, aluminum, iron and calcium. Oxides of magnesium, potassium, titanium and sulfur comprise approximately 4 percent of fly ash by weight (EPRI, 1995). Trace elemental composition of fly ash typically is approximately one percent by weight and may include arsenic, antimony, barium, boron, cadmium, chromium, copper, manganese, mercury, nickel, lead, selenium, silver, thallium, zinc, and other elements. For comparison, Figure 3-2 shows the elemental composition of fly ash and bottom ash compared with typical values for shale and volcanic ash. Table 3-2 shows the bulk composition of fly ash and bottom ash compared with typical values for soil and rock. In addition to these constituents, fly ash may contain unburned carbon. Bituminous coal ash typically yields slightly acidic to alkaline solutions (pH of 5 to 10) on contact with water. The geochemical factors controlling the reactions associated with leaching of ash are complex. Factors such as the chemical speciation of the constituent, solution pH, solution-to-solid ratio, and other factors control the chemical concentration of the resultant solution. Constituents that are held on the glassy surfaces of fly ash such as boron, arsenic, and selenium may initially leach more readily than other constituents. As noted in Table 3-2, aluminum, silicon, calcium, and iron represent the larger fractions of fly ash by weight. Calcium and iron may limit the release of arsenic by 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 3-4 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx forming calcium-arsenic precipitates. Formation of iron hydroxide compounds may also sequester arsenic and retard or prevent release of arsenic to the environment. Similar processes and reactions may affect other constituents of concern; however, certain constituents such as boron and sulfate will likely remain highly mobile. In addition to the variability that might be seen in the mineralogical composition of the ash, based on different coal types, different age of ash in the basin, etc., it is anticipated that the chemical environment of the ash basin varies over time, distance, and depth. EPRI (2010) reports that 64 samples of coal combustion products (including fly ash, bottom ash, and flue gas desulfurization residue) from 50 different power plants were subjected to United States Environmental Protection Agency (USEPA) Method 1311 Toxicity Characteristic Leaching Procedure (TCLP) leaching and no TCLP result exceeded the TCLP hazardous waste limit (EPRI, 2010). Figure 3-3 provides the results of that testing. 3.3 Site-Specific Coal Ash Data Source characterization was performed to identify the physical and chemical properties of the ash in the ash basin. The source characterization involved development of selected physical properties of ash, identifying the constituents found in ash, measuring concentrations of constituents present in the ash pore water, and performing laboratory analyses to estimate constituent concentrations resulting from the leaching process. The physical and chemical properties evaluated as part of this characterization will be used to better understand impacts to soil and groundwater from the source area and will also be utilized as part of groundwater model development in the CAP. Source characterization was performed through the completion of soil borings, installation of monitoring wells, and associated solid matrix and aqueous sample collection and analysis. Characterization of the ash basin was accomplished by completing five borings and installing nine monitoring wells in three phases. The first phase included borings that were installed using direct push technology (DPT) and continuous sample recovery. Each boring (AB-1, AB-2, AB-2, and AB-4) was advanced to the bottom of the basin. A second phase was conducted to collect samples of soil, saprolite, and bedrock beneath the ash, and install groundwater monitoring wells in various water-bearing horizons beneath and within the ash basin (ABMW wells). A third phase of work was completed to verify the lack of impact to deeper groundwater directly beneath the ash basin. This “data gap” effort resulted in the installation of deep monitoring well ABMW-2BRL. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 3-5 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx The initial assessment effort employing the DPT methods was undertaken to collect samples of ash and to determine the bottom of the deposited ash. A smaller, lighter DPT drilling rig was used so that conditions could be observed during transport and drilling activities and to determine if it was safe to move larger, heavier drilling equipment onto the ash basin. It was determined that the larger rotary sonic rigs could be employed if restricted to the so-called “ash harvest” area. This portion of the ash basin is where ash was removed from the basin and “stacked” for dewatering prior to transport for beneficial reuse. With the exception of ABMW-1, ash borings and wells were completed in the ash harvest area. Ash and soil samples were collected from each boring for physical and chemical testing in accordance with GAP Section 7.1.1 (SynTerra, 2014c) and as Site conditions allowed. Laboratory results of ash samples are presented in Appendix B, Table 4. The thickness of ash in the areas of investigation varied. Borings installed in the ash basin encountered ash from 13.5 feet to 66 feet in thickness. The contact between ash and underlying soils was distinct in each boring as seen in Photograph 3-1. Physical intrusion of ash into the underlying soils appeared to have been negligible. Physical Properties of Ash Physical properties (grain size, specific gravity, and moisture content) and mineralogy determinations were performed on samples from the ash basin. Physical properties were measured using ASTM methods and mineralogy was determined by X-ray diffraction (Appendix C). Bottom ash is generally characterized as a loose, poorly graded (fine- to coarse-grained) sand. Fly ash is generally characterized as a moderately dense silty fine sand or silt. Compared to soil, fly ash exhibits a lower specific gravity. Photograph 3-1. GeoProbe core from AB-2 (left) and AB-1 (right) showing one inch diameter sample in acetate tube. Soil (left of arrows) contact with saturated ash (right) is distinct with little vertical migration of ash. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 3-6 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx Mineralogy analyses on three ash samples indicate that the ash is predominantly mullite, with quartz, and one sample indicates 0.5 percent biotite content. Mullite is an aluminosilicate mineral (Al6Si2O13) that is rare in nature but common in artificial melts (Hurlbut, 1971). Presumably, the mullite formed from naturally occurring micas and clays in the coal-fired boiler. Quartz (SiO2) is the primary mineral in most natural sand deposits. Biotite (K(Mg,Fe) 3(AlSi3O10)(F,OH)2) is a main silicate found in coal (Table 3- 3; (Zhang, 2013). Three ash samples were tested by Energy Dispersive X-Ray Fluorescence (EDXRF) for metal oxides (Table 3-4) and a suite of elements (Table 3-5). The sample was comprised primarily of silicon dioxide (SiO2), aluminum oxide (Al2O3), and iron oxide (Fe2O3). Cerium, copper, tin, and zinc were the trace metals detected in highest concentrations in the sample. Chemical Properties of Ash Thirteen samples of ash were collected and analyzed for total constituent concentrations and total organic carbon (TOC). Concentrations of arsenic, chromium, cobalt, iron, and vanadium were reported above soil North Carolina Preliminary Soil Remediation Goals (PSRG) for Industrial Health and/or Protection of Groundwater (POG) for ash samples collected within the ash basin. Six samples were prepared using the USEPA Synthetic Precipitation Leaching Procedure (SPLP) (Appendix B, Table 6) and the leachate was analyzed. The SPLP was designed to more closely approximate leaching from a material by rainwater. SPLP leachate analytical results are compared to 2L and/or IMAC for reference purposes only. Those results do not represent groundwater samples; therefore, comparison to 2L and/or IMAC is not required. The SPLP analyses indicated leachate concentrations greater than 2L or IMAC in one or more samples for antimony, arsenic, chromium, iron, manganese, nitrate, thallium, and vanadium. Notably, boron was not detected in ash leachate above the 2L concentration. The SPLP is not intended to mimic complete leaching processes and results are not necessarily indicative of resultant concentrations in groundwater. Further, of the detected constituents in SPLP leachate data from the ash, cobalt, iron, manganese, and vanadium are prevalent in samples from background soil locations. Of these, cobalt and iron do not appear to readily leach from ash. Manganese and vanadium appear to leach equally from ash and natural soils. Chemistry of Ash Pore Water Pore water refers to water samples collected from wells installed within the ash basin and screened in the ash layer. Four pore water monitoring wells were completed in the ash basin (ABMW-1 through ABMW-4; Figure 2-8). Since installation of the wells in early 2015, the wells have been sampled seven times including the second quarter of 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 3-7 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx 2017 as part of the CAMA monitoring program. Due to unsafe accessibility after heavy precipitation events, ABMW-1 has been sampled only six times. Pore water analytical results are compared to 2L and/or IMAC for reference purposes only. The ash basin is a permitted wastewater system; therefore, comparison of pore water within the wastewater treatment residuals (ash) to 2B or 2L/IMAC is not required. Eleven analytes (antimony, arsenic, barium, boron, cobalt, iron, manganese, pH, total dissolved solids (TDS), thallium, and vanadium) were detected above the corresponding 2L or IMAC in one or more pore water samples (Appendix B, Table 1). Species of radium and uranium were detected in pore water from each well; however, only groundwater from ABMW-1 showed a concentration of total uranium which exceeded a comparison criterion. Concentrations of detected constituents in ash pore water have been relatively stable; although, decreases in iron and manganese have been noted in ABMW-2 and increases in iron and manganese have been noted in ABMW-4. Piper diagrams have been prepared for groundwater results from Mayo Plant monitoring wells and are discussed in more detail in Section 10.0. Relative redox conditions were determined using an Excel® workbook for identifying redox processes in ground water (Jurgens, McMahon, Chapelle, & Eberts, 2009). This workbook allows a standardized method to identify and describe the redox state of groundwater. Ash pore water from the basin is anoxic (ABMW-3, Fe(III) process), mixed (anoxic; ABMW-1, Fe(III)-SO4 process), and mixed (oxic/anoxic; ABMW-2 and ABMW-4, O2-Fe(III)-SO4 process). Ash pore water, as well as groundwater beneath the ash basin, is predominantly characterized as calcium-bicarbonate type water, typical of shallow fresh groundwater but atypical of coal ash leachate. For comparative purposes, a 2006 EPRI study of 40 ash leachate water samples collected from 20 different coal ash landfills and impoundments characterized bituminous coal ash leachate as calcium-magnesium-sulfate water type and subbituminous coal ash leachate as sodium-calcium-sulfate water type. Most of the ash pore water and groundwater within the ash basin exhibits neutral or slightly basic pH as contrasted with the slightly acidic groundwater present in background, sidegradient, and many downgradient wells. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 4-1 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx 4.0 RECEPTOR INFORMATION Section §130A-309.201(13) of the CAMA defines receptor as “any human, plant, animal, or structure which is, or has the potential to be, affected by the release or migration of contaminants. Any well constructed for the purpose of monitoring groundwater and contaminant concentrations shall not be considered a receptor.” In accordance with the NORR CSA guidance, receptors cited in this section refer to public and private water supply wells (including irrigation wells and unused wells) and surface water features. Additional receptors (described in Section 12.0) were evaluated as part of the risk assessment related to the CSA effort. The NORR CSA receptor survey guidance requirements include listing and depicting water supply wells, public or private, including irrigation wells, and unused wells (other than those that have been properly abandoned in accordance with 15A NCAC 2C .0100) within a minimum of 1,500 feet of the known extent of contamination. In NCDEQ’s June 2015 response to Duke Energy’s proposed adjustments to the CSA guidelines, NCDEQ DWR acknowledged the difficulty with determining the known extent of contamination at this time. DWR stated that it expected all drinking water wells located 2,640 feet (0.5-mile) downgradient from the established compliance boundary to be documented in the CSA reports as specified in the CAMA requirements. The approach to the receptor survey in this CSA includes listing and depicting all water supply wells (public or private, including irrigation wells, and unused wells) within a 0.5-mile radius of the ash basin compliance boundary (Appendix D). Properties located within a 0.5-mile radius of the ash basin compliance boundary include residences located to the south and upgradient of the Site, centered around Mullins Lane; residences located northwest and upgradient of the Site, on the south/North Carolina side of the North Carolina/Virginia state line; and residences located northwest and upgradient of the Site, on the north/Virginia side of the state line. Unoccupied rural properties are located north of the Site and downgradient of the ash basin (Figure 4-1). A municipal water line is present along US Highway 501 (Boston Road) toward the south of the Plant. The water line does not extend north of the Plant entrance road. No municipal water lines serve the area north of the Site (along Mayo Lake Road). The NORR CSA guidance requires that subsurface utilities be mapped within 1,500 feet of the known extent of contamination in order to evaluate the potential for preferential pathways. Identification of piping near and around the ash basin was conducted by Stantec in 2014 and 2015 and utilities around the Site were also included on a 2015 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 4-2 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx topographic map by WSP USA, Inc. (Appendix E). Culverts, pipes, and miscellaneous subsurface utilities were also located on the Mayo Site. Due to the isolation of the ash basin from the Plant area, subsurface utilities in the Plant area are not expected to be major contaminant flow pathways. The depth to groundwater below the majority of the ash basin is much greater than would be anticipated for installation of subsurface utilities; therefore, the likelihood of underground utilities being preferential pathways, other than at the dam seepage structures, is not anticipated. 4.1 Summary of Receptor Survey Activities Surveys to identify potential receptors including public and private water supply wells (including irrigation wells and unused or abandoned wells), and surface water features within a 0.5-mile radius of the Mayo ash basin compliance boundary have been reported to NCDEQ: Drinking Water Well and Receptor Survey – Mayo Steam Electric Plant (SynTerra, 2014a) Supplement to Drinking Water Well and Receptor Survey – Mayo Steam Electric Plant (SynTerra, 2014b) Update to Drinking Water Well and Receptor Survey – Mayo Steam Electric Plant (SynTerra, 2016c) These reports are included as Appendix D. The Drinking Water Well and Receptor Survey – Mayo Steam Electric (SynTerra, 2014a) included results of a review of publicly available data from NCDEQ Division of Environmental Health, NC OneMap GeoSpatial Portal, DWR Source Water Assessment Program (SWAP) online database, county geographic information system, Environmental Data Resources, Inc. (EDR) records review, the USGS National Hydrography Dataset, as well as a vehicular survey along public roads located within 0.5-mile radius of the ash basin compliance boundary. The Supplement to Drinking Water Well and Receptor Survey- Mayo Steam Electric Plant (SynTerra, 2014b) supplemented the initial report with additional information obtained from questionnaires completed by property owners who own property within the 0.5- mile radius of the ash basin compliance boundary. The report included a sufficiently scaled map showing the coal ash facility location, the boundary of the Site, the waste and compliance boundaries, all monitoring wells listed in the NPDES permit and the approximate location of identified water supply wells. A table presented available 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 4-3 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx information about identified wells including the owner's name, address of the well with parcel number, construction and usage data, and the approximate distance from the compliance boundary. The Update to Drinking Water Well and Receptor Survey – Mayo Steam Electric Plant (SynTerra, 2016c) included review of the most recent available state, county and other resources and an additional field reconnaissance to observe potential water supply wells near a 0.5 mile radius of the established Mayo Plant ash basin compliance boundary. No additional wells were identified near the survey areas that were not reported in the 2014 surveys. Consistent with findings from previous surveys, no public or private drinking water wells or wellhead protection areas are located downgradient of the ash basin. 4.2 Summary of Receptor Survey Findings The City of Roxboro provides and maintains a municipal water supply line that extends north from Roxboro along US Highway 501. The line supplies potable water to the Mayo Plant. The line terminates on US Highway 501 at the Mayo Plant. No public or private drinking water wells or wellhead protection areas were found to be located downgradient of the ash basin. This finding was supported by field observations and a review of public records. Based on the known groundwater flow direction, none of the wells identified in the water well survey are located downgradient of the ash basin. The location and relevant information pertaining to suspected water wells located upgradient of the facility, within 0.5 miles of the ash basin compliance boundary, were included in the survey reports as required by the NORR. As required by G.S. 130A-309.211(c1) of House Bill 630 (HB630), Duke Energy evaluated the feasibility and costs of providing a permanent replacement water supply to eligible households. Households were eligible if any portion of a parcel of land crossed the 0.5- mile compliance line described in House Bill 630 and if the household currently was using well water or bottled water (under Duke Energy’s bottled water program) as the drinking water source. Undeveloped parcels were identified but were not considered “eligible” because groundwater wells are not currently utilized as a drinking source. A Potable Water Programmatic Evaluation (Dewberry, 2016); (Appendix D) was conducted. That evaluation included a survey of eligible households, a preliminary engineering evaluation, a cost estimate, and a schedule. The evaluation report also included a listing of, and relevant information about, households and properties within the survey area, 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 4-4 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx as well as, maps depicting property locations, including those properties for which a replacement water supply will be provided. Public Water Supply Wells 4.2.1 An EDR Report for the nearby Louisiana Pacific Corporation site west of US Highway 501 indicated that Bethel Hill Baptist Church, located approximately 0.5 miles south of the Site at 201 Old US Hwy 501 (Roxboro, North Carolina), maintains a public water supply provided by a groundwater well. Private Water Supply Wells 4.2.2 The fractured bedrock aquifers in the north-central Piedmont, including in the rural areas surrounding the Mayo Plant, are commonly used for water supply purposes. Occupied properties southwest and northwest of the Site are not serviced by the municipal water supply line. Drinking water is obtained from private groundwater wells by residents of the following: residences on or near Mullins Lane located just south of the Mayo Plant; one business and one residence northwest of Boston Road and the Mayo Plant just south of the state line; and several residences along the east and west sides of US Highway 501 (Huell Matthews Highway) in Virginia, just north of the state line (Figure 4-1). Several efforts have been made to locate and document the presence of and information related to private water supply wells in the vicinity of the Mayo Plant. The September 2014 Drinking Water Well and Receptor Survey report indicated that no public or private drinking water wells or wellhead protection areas were located within the 0.5-mile off-set from the compliance boundary; however, private water supply wells have been identified within or in close proximity to the 0.5-mile off-set. SynTerra’s November 2014 report supplemented the initial report with information obtained from questionnaires sent to owners of property within the 0.5-mile radius of the compliance boundary. The questionnaires were designed to collect information regarding whether a water well or spring is present on the property, its use, and whether the property is serviced by a municipal water supply. If a well is present, the property owner was asked to provide information regarding the well location and construction information. The results from the questionnaires indicated that as many as 22 wells might be located within 0.5 miles of the compliance boundary for the Site (reported wells, observed wells, and possible wells). The November 2014 and September 2016 update reports included a sufficiently scaled map showing the ash basin location, the facility property boundary, the 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 4-5 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx waste and compliance boundaries, all monitoring wells, and the approximate location of identified water supply wells. A table presented available information related to identified wells, including: the owner's name, the address of the well location with parcel number, construction and usage data, and the approximate distance of the well from the compliance boundary. There are three inactive water supply wells on the Site (informally designated as DEP-1, DEP-2, and DEP-3; Figure 2-8). These wells were likely drilled in the late 1970s or early 1980s (and possibly 1990s in the case of DEP-1) to supply water to the Plant during construction and early start-up operations (DEP-2 and DEP-3). DEP-1 was drilled to provide potable water to the Plant picnic area (no longer in use). An evaluation of the wells was conducted during Site assessment activities. Pumps and piping were removed, and the depths of the well and static water levels were measured. A downhole camera was used in an attempt to video-log the entire well to obtain well-specific data. The video logging effort was only partly successful because the wells have been idle for many years and a significant biofilm has formed and coated the walls of the borehole. This biofilm causes very turbid and murky water. In spite of the poor condition of the well water, relevant information about the wells was obtained (table below). It appears, based on the nature of the rock viewed with the downhole camera, that the wells were drilled using a pneumatic air hammer. It is noteworthy that the wells have been out of service for a number of years, even decades; therefore, they are not currently influencing groundwater flow. WELL DESIGNATION TOTAL WELL DEPTH (feet bgs) DEPTH OF SURFACE CASING (feet bgs) WATER LEVEL (feet bgs; July 2015) WELL INFORMATION DEP-1 130 36 10.40 Casing: 6-inch steel; Pump: 1.5 HP (household type); Discharge Pipe: 1-inch (black poly) DEP-2 238 21 15.65 Casing: 6-inch steel; Pump: 5 HP (Diamond; installed 10/31/1995); Pump Set Depth: 215 ft bgs; Discharge Pipe: 1.25-inch (galvanized) DEP-3 250 21 21.23 Casing: 6-inch steel; Pump: 5 HP (Gould); Pump Set Depth: 227 ft bgs; Discharge Pipe: 2-inch (galvanized) 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 4-6 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx 4.3 Private Water Well Sampling NCDEQ coordinated sampling of six private water supply wells within a half-mile radius of the ash basin compliance boundary from March to May in 2015. Water samples were collected from three of the six wells identified by property owners. Only one well owner reported a well. Duke Energy collected samples from eight additional private water supply wells in 2017. The three original wells sampled by NCDEQ were not resampled by Duke Energy. Tabulated results, provided by Duke Energy, for the NCDEQ and Duke Energy sampling efforts along with exceedances of 2L Standards, IMACs, and/or other reporting limits are included in Appendix B, Table 1. The locations of the sampled wells and summary analytical results are included on Figure 4-2. A review of the analytical data for the private water supply wells indicated several constituents were detected above 2L or IMACs including pH (two wells), iron (three wells), lead (one well), manganese (three wells), and vanadium (eight wells). Concentrations of analyzed constituents exceeded the respective Provisional Background Threshold Values (PBTV) for a number of private water supply wells (high turbidity data/values are excluded) including: Alkalinity (1 well) Barium (2 wells) Calcium (1 well) Chromium (hexavalent) (4 wells) Copper (10 wells) Lead (7 wells) Manganese (1 well) Molybdenum (1 well) Potassium (2 wells) Selenium (1 well) Sodium (1 wells) Strontium (2 wells) Sulfate (2 wells) TDS (2 wells) 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 4-7 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx Vanadium (1 well) Zinc (7 wells) The exceedances of PBTVs in private water wells were further evaluated. First, the PBTVs have been developed using groundwater data from a set of three background wells from a geographically limited area, all located on the Mayo Plant. These wells are located within about one-tenth of square mile of each other. The geochemical data from these wells may not be representative across the broader area encompassed by the 11 private water supply wells (spread across approximately 1.5 square miles). For example, if the Mayo dataset were compared with the PBTVs for the nearby and similar Roxboro Plant (e.g., similar geologic setting), only 10 of the 16 constituents listed above would exceed a background value in groundwater. Second, well construction may influence analytical results. For example, galvanized pipe could yield high zinc concentrations and brass components in wells pumps and valves can be a source of lead. Information concerning well construction and piping materials is important to have before attributing detections of ash-related constituents solely to the geochemistry of the groundwater. Third, there is very limited information available about the wells (e.g., date of installation, drilling method, well depth, casing length, pump set depth, etc.). Many private water supply wells in this part of the Piedmont are open-hole rock wells. A shallow surface casing is installed and then the well is drilled to a depth that may be as shallow as 40 or 50 feet or as deep as several hundred feet. When a groundwater sample is collected, it is unknown from what part of the bedrock aquifer the groundwater is drawn. Groundwater geochemistry in fractured bedrock aquifers can be quite variable. A fourth reason for considering the apparent exceedances of PBTVs in groundwater is that, as previously described, private water wells in bedrock are typically installed as open-hole wells. Care must be taken when comparing geochemical data from these wells and comparing them to background concentrations derived from carefully drilled and installed groundwater monitoring wells with machine-slotted wells screens, proper filter pack installation, proper well development, and specific sample collection procedures employed. Fifth, groundwater flow in the area around Mayo is consistent with the model of groundwater flow in the Piedmont as described in Section 5.2. This conceptual model of groundwater flow (LeGrand, 1988) (LeGrand, 1989) describes each distinct Piedmont surface drainage basin as similar to adjacent basins with the conditions generally repetitive from basin to basin. Within a basin, movement of groundwater is generally 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 4-8 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx restricted to the area extending from the drainage divide to a perennial stream (Slope- Aquifer System). Each distinct basin and slope-aquifer system may limit the area of influence of wells. In the case of the private wells located upgradient of the Mayo Plant, the wells are situated in distinct drainage basins/slope-aquifer systems separate from the Plant area and the ash basin. Further, groundwater flow from the source (ash basin) is to the north-northeast and the closest private wells are situated to the north- northwest, northwest, and south-southeast of the ash basin. Finally, the geochemical signature of groundwater from the private wells was compared with the signature of groundwater from the source area using Piper diagrams (a graphical representation of major water chemistry using two ternary plots and a diamond plot showing the relative percentage of major cations and major anions in a sample). The geochemical nature of groundwater from the sampled private wells is very different from ash pore water and from groundwater beneath the basin (discussed in Section 10.0). Data from five of the wells was plotted because data with a charge balance greater than 10 percent were omitted. Four of the wells (MY-03, MY-1000, MY- 1003, and MY-2003) plot consistently with deep bedrock background well MW-13BR and typical, deep groundwater geochemical types. MY-1002 results suggest a sodium bicarbonate water type, indicative of ion exchange, possibly indicating that water treatment is associated with that well. 4.4 Numerical Well Capture Zone Analysis In December 2015, a numerical capture zone analysis for the Mayo Site was conducted to evaluate potential impact of upgradient water supply pumping wells. The analysis employed MODPATH to interface with the MODFLOW flow model. MODPATH is a “particle tracking” model that traces groundwater flow lines from a starting position. MODPATH was used to trace groundwater flow lines around pumping wells to indicate where the water being pumped from the well originates (i.e., well capture zone analysis). The analysis for Mayo indicates that well capture zones from wells located to the northwest and southeast of the Mayo Plant are limited to the immediate vicinity of the well head and do not extend toward the ash basin. None of the particle tracks originating in the ash basin moved into the well capture zones (Figure 4-3). 4.5 Surface Water Receptors The Site is located in the Roanoke River Basin. Although the Site is located near Mayo Lake, groundwater influenced by the ash basin flows toward Crutchfield Branch, a small stream located north of the ash basin and near the northern Site property. There is no surface water intake in Crutchfield Branch. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 5-1 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx 5.0 REGIONAL GEOLOGY AND HYDROGEOLOGY North Carolina is divided into distinct regions by portions of three physiographic provinces: the Atlantic Coastal Plain, the Piedmont, and the Blue Ridge. The Mayo Plant is situated in the Piedmont physiographic province of north-central North Carolina. The Piedmont is generally characterized by mature, well-rounded hills and rolling ridges cut by small streams and drainages. However, in areas with thinner regolith, like in the immediate vicinity of the Mayo Plant, the relief is more rugged with incised streams that occur as the rate of subsurface weathering fails to keep pace with the rate of erosion. The Piedmont in North Carolina is several hundred feet higher than in neighboring South Carolina and Virginia due to the Cape Fear Arch, an uplift feature that trends roughly along the Cape Fear River and continues through the Piedmont into the Appalachian Mountains (Rogers, 2006). The resulting geomorphology results in river flow to the north or south instead of east (Rogers, 2006). Elevations in the area of the Mayo Plant range from 600 feet in the extreme southwest portion of the Plant property (near the Plant entrance along Boston Road) to 360 feet in the Crutchfield Branch stream area (on the north side of the Plant). The following sections contain a synopsis of geologic and hydrogeologic characteristics for the area. This section does not provide an exhaustive list or summary of the many important geologic research efforts that have been published on the region. This section provides summary information from research. 5.1 Regional Geology The Plant is located near the contact between two regional zones of metamorphosed rocks: the Carolina Slate Belt (often referred to as Carolina terrane) on the east and the Charlotte Belt (or Charlotte terrane) to the west (Figure 5-1). The majority of the Mayo Site, including the largest portion of the ash basin and Mayo Lake, is situated in the Carolina terrane (Dicken, Nicholson, Horton, Foose, & Mueller, 2007). The Carolina terrane includes volcanic and sedimentary rocks metamorphosed to lower greenschist facies (Butler & Secor, 1991). The metamorphic rocks have been intruded by coarse- grained granitic rocks and have been subjected to regional structural deformation (Rogers, 2006). For more than a century, the character and genesis of the rocks within these regional metamorphic belts has been the subject of intense study and efforts to describe the mineral resources of the area and the geologic character of the area in tectonic, structural, and litho-stratigraphic terms. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 5-2 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx In a work co-published by the Virginia Geological Survey and the North Carolina Geological and Economic Survey in 1917, Laney traced the history of investigations in and around the so-called Virgilina Ore District from the early 1700s, through the 19th century, and into the early 20th century (Laney, 1917). The Virgilina District was an important copper ore producing area located within the Carolina Slate Belt of North Carolina and Virginia. Laney (1917) provides detailed petrographic, mineralogical, stratigraphic, and structural geologic descriptions of the rocks and geologic formations observed and mapped in the Virgilina District noting the broad array of igneous and metamorphic rock types present. Carpenter (1976) provided an update on North Carolina’s metallic mineral industry associated with Carolina Slate Belt rocks as well as a descriptive summary of the variety of rocks encountered. Carpenter made note of the relatively “higher rank metamorphic rocks (that) occupy the northeast corner” of Person County. Carpenter also described the geology of the mines in Person County, and his description of the geology encountered at a mine located several miles east of the present-day location of the Mayo Plant noted the presence of chlorite-sericite schist and phyllite with “stringers of epidote, calcite, and quartz” as well as evidence of shearing and faulting within the schist (Carpenter, 1976). These descriptions are consistent with rocks encountered during field work (Section 6.0 Site Geology and Hydrogeology). In 1973, Glover and Sinha (1973) published research results on the evolution of the rocks of the central Piedmont of North Carolina and Virginia (Glover & Sinha, 1973). Their extensive mapping work indicated that the present-day Mayo Plant is located near/along a contact between felsic and intermediate volcanic rocks (Hyco Formation) and “tuffaceous epiclastic rocks and reworked tuffs” that commonly present as phyllitic siltstone, phyllite, and other minor occurrences of conglomerate and pyroclastics (Aaron Formation). The authors set forth their hypothesis for a major deformational event called the Virgilina deformation that produced major structures in the central Piedmont. Glover and Charles Harris revisited the Virgilina deformation theory with a summary of the hypothesis and additional detail and geologic mapping updates concerning the stratigraphic and petrologic character of the rocks of the north-central Piedmont in the Roxboro, North Carolina/Virgilina, Virginia area (Harris & Glover, 1985). The Geologic Map of North Carolina (1985) describes the area of the Mayo Plant as underlain by felsic meta-volcanic rock interbedded with mafic and intermediate volcanic rock, meta-argillite and meta-mudstone (NCDNRCD, 1985). In 1991, Butler and Secor (1991) presented an exhaustive summary of geologic research on the rocks of the Central Piedmont of North Carolina with a subsection pertaining to 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 5-3 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx the genesis, stratigraphy, rock associations, rock ages, and structural relationships of the northern Carolina slate belt. Researchers in the 1990s began to refer to the slate belt as Carolina “terrane” or the Carolina “zone” in terms of the tectonostratigraphic relationships between the rock assemblages (Butler & Secor, 1991). The Carolina zone was described as “a lower greenschist to amphibolite sequence of felsic and mafic volcanics and metasedimentary rocks intruded by felsic to mafic plutonic rocks of various ages” (Wilkins, Shell, & Hibbard, 1995). Foliations within the metamorphic rocks of the Carolina zone are described as steeply dipping along upright folds. These rock types and rock structures are contrasted with the higher metamorphic grades and generally gently dipping foliations and recumbent folds in the Piedmont zone (Wilkins, Shell, & Hibbard, 1995). Hibbard and his co-authors (1998) describe the north-northeast trending Hyco shear zone as a terrane boundary separating the Piedmont (and locally, the Charlotte and Milton terranes) from the Carolina terrane” (Hibbard, Shell, Bradley, & Wortman, 1998). The Hyco shear zone extends to the east about 8 kilometers into Carolina terrane rocks – and into the vicinity of the Mayo Plant – and is defined as the hanging wall of the fault zone. Confirming the earlier mapping of Glover and Sinha (1973) and Harris and Glover (1985), the rocks of the eastern edge of the Hyco shear zone occur along the western edge of the Mayo Plant area and are described as “dominantly greenschist facies felsic volcanic and volcaniclastic rocks with subordinate intermediate and mafic components” (Glover & Sinha, 1973); (Harris & Glover, 1985). The authors provide a detailed analysis of the structural nature of the Hyco shear zone and state their position concerning the timing and large-scale regional tectonic implications of their work (Hibbard, Shell, Bradley, & Wortman, 1998). Since 2000, additional research and contributions to the relevant scientific literature on the Carolina terrane have continued to refine a model of the complicated tectonic, stratigraphic, and lithographic character and interrelationships of the region, notably Hibbard, Stoddard, Secor, and Dennis (2002); Bowman (2010); Pollock, Hibbard, and Sylvester (2010); and Bowman, Hibbard, and Miller (2013). 5.2 Regional Hydrogeology The upper portions of rocks in the Piedmont are typically fractured and weathered and are covered with unconsolidated material known as regolith. The regolith includes residual soil and saprolite zones and, where present, alluvium. Saprolite is formed by in-situ chemical weathering of bedrock. It is typically composed of clay and coarser granular material and reflects the texture and structure of the parent rock. For example, the weathering products of granitic rocks are quartz-rich and sandy textured. Rocks 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 5-4 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx poor in quartz but rich in feldspar and ferro-magnesium minerals form a more clayey saprolite. The degree of weathering decreases with depth, and partially weathered rock (PWR) is commonly present near the top of the bedrock surface. The transition zone from the regolith and the PWR and competent bedrock is often gradational and difficult to differentiate. Groundwater flow systems in the Piedmont are comprised of two interconnected hydrogeologic units: (1) residual soil/saprolite and weathered fractured rock (regolith and PWR) overlying (2) fractured crystalline bedrock (Heath, 1980); (Harned & Daniel, 1992). The regolith layer is a weathered and structureless residual soil that occurs near ground surface with the degree of weathering decreasing with depth. Residual soil grades into saprolite. Beneath the saprolite, partially weathered/fractured bedrock occurs with depth until competent bedrock is encountered. This mantle of residual soil, saprolite, and weathered/fractured rock (transition zone) is a hydrogeologic unit that covers and crosses various types of rock (LeGrand, 1988). This layer serves as the principal storage reservoir and provides a granular medium through which the recharge and discharge of water from the underlying fractured rock occurs (Harned & Daniel, 1992). A transition zone at the base of the regolith is present in many areas of the Piedmont. The zone consists of partially weathered/fractured bedrock and lesser amounts of saprolite that grades into competent bedrock and has been described as “being the most permeable part of the system, even slightly more permeable than the soil zone” (Harned & Daniel, 1992). The zone thins and thickens within short distances and its boundaries may be difficult to distinguish. Where present, the zone may serve as a conduit of rapid flow and transport of impacted groundwater (Harned & Daniel, 1992). Daniel and Dahlen (2002) provide a summary of the nature and occurrence of groundwater in fractured rock. Within the fractured crystalline bedrock, fracture apertures, connectivity, etc. control groundwater movement and storage capacity. The bedrock is broken and displaced by faults and shear zones, some of which extend for miles. Joints, rock breaks without accompanying displacement, are common, and the joints typically occur in groups oriented in preferred directions. Weathering and erosion have resulted in fracturing in the form of stress-relief fractures, as well as expansion of existing fractures, and it is through these fractures that groundwater flows. Planes and bedding of metamorphic foliation, as well as breaks and folds in these rocks, are areas of higher permeability (Daniel & Dahlen, 2002). 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 5-5 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx LeGrand’s (1988; 1989) conceptual model of the groundwater setting in the Piedmont incorporates the two-medium system described above into a single feature that is useful for the description of groundwater conditions. That feature is the surface drainage basin that contains a perennial stream (LeGrand, 1988). In general terms, each surface drainage basin is similar to adjacent basins and the conditions are generally repetitive from basin to basin. Within a basin, movement of groundwater is generally restricted to the area extending from the drainage divides to a perennial stream (Slope-Aquifer System; (LeGrand, 1988); (LeGrand, 1989); Figure 5-2). Freeze and Witherspoon’s (1967) model for regional groundwater flow centers on a large regional discharge area that will receive water from a groundwater basin except for shallow discharges into smaller perennial streams located closer to sub-regional recharge areas (Freeze & Witherspoon, 1967). Shallow groundwater near perennial streams will discharge into that stream. The crests of water table undulations represent natural groundwater divides within a slope-aquifer system and may limit the area of influence of wells or contaminant plumes located within their boundaries depending on the depth of the impacted groundwater. The concave topographic areas between the topographic divides may be considered as flow compartments that are open-ended down slope. Therefore, in most cases in the Piedmont, the groundwater system is a two-medium system (LeGrand, 1988) restricted to the local drainage basin. Groundwater within the area exists under unconfined (water table) conditions within the saprolite, PWR/transition zone, and in the fractures and joints of the underlying bedrock. The water table and bedrock aquifers are often interconnected. Typically, the residual soil/saprolite is partially saturated and the water table fluctuates within it. The saprolite and PWR/transition zone acts as a reservoir for water supply to the fractures and joints in the underlying bedrock. Groundwater recharge in the Piedmont is derived entirely from infiltration of local precipitation. Groundwater recharge occurs in areas of higher topography (i.e., hilltops) and groundwater discharge occurs in lowland areas bordering surface water bodies, wetlands, and floodplains (LeGrand, 2004). Average annual precipitation in the Piedmont ranges from 40 inches to 50 inches with a minimum of about 30 inches and a maximum of about 80 inches. Mean annual recharge in the Piedmont ranges from about 4 inches to under 10 inches (Daniel & Dahlen, 2002). 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 6-1 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx 6.0 SITE GEOLOGY AND HYDROGEOLOGY Geology beneath the Mayo Plant can be classified into three units. Regolith (surficial soils, fill and reworked soil, alluvium along the Crutchfield Branch stream valley, and saprolite) is the shallowest geologic unit. Saprolite is mostly thin (ranging from nonexistent to around 25 feet deep) and almost entirely unsaturated. This generalization is not consistent for the southern, upland parts of the Site where a thick, saturated saprolite zone is present (e.g., MW-12 well pair) nor for certain locations beneath the ash basin. The thin to nonexistent saprolite zone across the central and northern portion of the Site is due to extensive excavation and reworking of surficial materials during Plant construction. A transition zone of partially weathered rock underlies the regolith (where present, the saprolite is the lowest portion of the regolith) and is generally continuous throughout the Mayo Plant area. However, the transition zone at the Mayo Site is comprised mostly of partially weathered rock that is gradational between saprolite and competent bedrock. The change from partially weathered rock to the third unit, competent bedrock, is subjective, and at the Mayo Site, is defined by subtle changes in weathering, secondary staining and mineralization, core recovery, and the degree of fracturing in the rock. Typically, mildly productive fractures (providing water to wells) were observed within the top 50 feet to 75 feet of competent rock. In general, three hydrogeologic units or zones of groundwater flow can be described for the Site. The zone closest to the surface is the shallow or surficial flow zone encompassing saturated conditions, where present, in the residual soil, saprolite, or alluvium beneath the Site. A transition zone, encountered below the surficial zone and above the bedrock, is characterized primarily by partially weathered rock of variable thickness. The transition zone is not consistently saturated across the Site. The bedrock flow zone occurs below the transition zone and is characterized by the storage and transmission of groundwater in water-bearing fractures. Site investigations included installation of soil borings, collection of soil and rock cores, groundwater monitoring wells, borings in and through the ash basin, and installation of wells for the sampling of ash pore water. Physical and chemical properties of soil samples collected from the borings and wells are presented in Tables 6-1 and 6-2, respectively. The analytical methods used with solid and aqueous samples are presented in Table 6-3 and Table 6-4. Table 2-1 summarizes the well construction data for CAMA-related wells and piezometers at the Site. Strategic locations for anchoring flow path transects were selected. Boring logs for CAMA-related monitoring 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 6-2 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx installations are included in Appendix F. Primary technical objectives for the sampling locations included: the development of additional background data on groundwater quality; the determination of horizontal and vertical extent of impact to soil and groundwater; and the establishment of perimeter boundary conditions for groundwater modeling that will be used to develop a CAP. The Mayo Plant ash basin occupies the former stream valley of Crutchfield Branch. The basin acts as an elongated bowl-like feature with groundwater flowing to the basin from all sides, except from the northeast, which is the discharge side of the basin. Groundwater flows north-northeast from the ash basin into the small valley formed by Crutchfield Branch. Crutchfield Branch flows north off of the Site property into Virginia. Mayo Lake affects groundwater on the east side of the Plant and acts as a groundwater discharge area. Groundwater flows from the highest topographic portion of the Site (near the Plant entrance road) to the north and northeast. US Highway 501 crests a ridge just west of the western edge of the ash basin (near MW-13BR) and bisects groundwater flow to the west and east. 6.1 Site Geology The regional geologic setting for the Mayo Plant is described in Section 5.0. A geologic map for the Mayo Plant area is included as Figure 6-1. The subsurface at the Site is composed of regolith (including residual soils, fill and reworked soils, alluvium, and saprolite), transition zone, and bedrock. Each zone was not encountered at every boring location. Subsurface conditions varied with topography, parent rock, and Site infrastructure. Alluvium was observed at two locations along Crutchfield Branch (MW-16S and SB-7), with a maximum thickness of 7.5 feet. The dominant rock types observed were granitoid gneiss and mica gneiss with minor mica schist and phyllite. The metavolcanic rocks observed were largely felsic with some mafic metavolcanics and meta-argillites interbedded. In general, metamorphic grading increases to the southwest and west of Mayo Plant, with plutonic volcanic bodies observed to the west. Soil Classification 6.1.1 Regolith was encountered from a depth range of a few inches to 66 feet bgs at the Site. A distinct organic soil horizon was rarely seen and, where present, was often only a few inches thick. Grain size analyses of soils indicate that the major soil texture at the Site is sandy loam, with layers of loamy sand, loam, and clay. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 6-3 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx Mineralogical analyses indicate the presence of quartz, feldspars, amphibole, chlorite, and a range of clay minerals (chiefly illite) in Site soils. Mineralogical analyses indicate that clay minerals comprise the bulk portion of Site soils, followed by quartz, feldspars and amphiboles in order of decreasing abundance. The mineralogical partitioning varied from one location to another. The regolith was dominated by saprolite, the in-situ weathering product of parent rock, and the mineralogical composition of the saprolite varied with subsurface lithology. At MW-12D, a background well with 66 feet of saprolitic regolith, the mineral assemblage consisted of predominantly clay and felsic minerals with 52.5 percent illite, 14.6 percent kaolinite, 22.6 percent quartz, and 10.0 percent feldspars. In contrast, the mineralogical data at ABMW-2BR, in the saprolitic regolith below the ash basin shows the presence of significantly more mafic components with: 33.5 percent quartz, 32.8 percent amphibole, 15.8 percent feldspars, and 17.8 percent clay minerals (smectite and mullite). The composition differences are evidence of the change in parent rock from largely mica gneiss on the western portion of the Site and more mafic metavolcanics and meta-argillites to the east. Saprolite often contains relict structure from the parent rock, retaining directional properties and permeability (Daniel & Dahlen, 2002). Photograph 6-1 illustrates the relict structure observed in saprolite from the MW-12S/12D background location. Rock Lithology 6.1.2 The Mayo Plant is located near the contact between the Carolina and Charlotte (or Milton) terranes, within the Hyco Shear Zone (Hibbard, Shell, Bradley, & Wortman, 1998). Charlotte terrane rock formations consisted of felsic mica and granitoid gneiss interbedded with hornblende gneiss and phyllites. The Carolina terrane formations consist of metamorphosed dacitic to rhyolitic flows and tuffs interbedded with mafic and intermediate metavolcanic rock, meta-argillite, and meta-mudstone. The contact has a general northeast–southwest trend with the Charlotte terrane on the west and the Carolina terrane on the East. The Photograph 6-1: MW-12D (39.0 ft below ground surface (BGS)) Saprolite showing relict structure. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 6-4 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx metamorphosed rocks have been intruded by coarse-grained granitic rocks and have been subject to regional structural deformation (Rogers, 2006). On the west side of the terrane contact, gneiss consisted of fine to coarse mica gneiss, mica schist, and granitoid layers. Hornblende gneiss in the region is found in close association with this felsic gneiss and schist as narrow and irregular areas and dike-like bands (Laney, 1917). An irregular hornblende gneiss zone within the felsic mica gneiss was observed at the MW-5BR location as illustrated in Photograph 6-2. East of the terrane contact, the subsurface lithology consists of granitoid gneiss, light gray, dark gray and distinctively green meta-argillite and metavolcanic rock zones. Chemical analysis indicates an increase in chlorite, biotite and muscovite in rock samples taken from the northeastern portion of the Site. Chlorites and micas are minerals found in the greenschist facies, and according to Butler and Secor (1991), the Carolina terrane consists of volcanic and sedimentary rocks metamorphosed to lower greenschist facies. Photograph 6-3 illustrates a zone of metavolcanic rock of the greenschist facies encountered at MW-7BR, northeast of the Plant. At the terrane contact, shear stress is evident by mylonitic texture identified at several boring locations. Photograph 6-4 exemplifies the mylonitic texture observed in some Mayo Plant bedrock. Photograph 6-2: MW-5BR (42.0-44.5 ft BGS). Sharp contact between felsic and mafic gneiss. Photograph 6-3: MW-7BR (57.0-67.0 ft BGS) Greenschist facies rock to 65.7 ft BGS sharp contact with granitic gneiss. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 6-5 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx Structural Geology 6.1.3 Bedrock underlying the Mayo Plant is a combination of Neoproterozoic- Early Paleozoic, metaigneous-dominated Charlotte (or Milton) and Carolina terranes. The metamorphic bedrock is associated with the Hyco Shear Zone, which is an east-northeast to north-northeast trending structure that separates the infrastructural Charlotte terrane from the suprastructural Carolina terrane (Figure 5-1). The shear zone contains Charlotte terrane rocks in the footwall to the west and Carolina terrane rocks in the hangingwall to the east (Hibbard, Shell, Bradley, & Wortman, 1998). The younger Charlotte terrane contains plutonic rocks that intrude a suite of metaigneous rocks of the amphibolite facies with minor phyllite, mica schist, and quartzite (Hibbard et al., 1998, 2002). According to Dennis and Shervais (1991, 1996), “On the basis of geochemical studies, the metamafic complexes have been interpreted as arising from suprasubduction zone magmas related to an episode of arc-lifting.” The older Carolina terrane is generally lower-grade metaigneous and associated metasedimentary rocks, of the greenschist to amphibolite facies. Four sequences comprise the bulk of the Carolina terrane, the Virgilina sequence, the Albemarle sequence, the South Carolina sequence, and the Cary sequence. The Virgilina sequence is the oldest (633 Ma to 612 Ma) of the four. It underlies the portion of the Carolina terrane along the Hyco Shear Zone and includes the Hyco, Aaron, and Virgilina Formations (Wortman, Samson, & Hibbard, 2000); (Hibbard, Stoddard, Secor, & Dennis, 2002). The Hyco Formation includes a thick felsic- intermediate metavolcanic base overlain by metaclastic turbidites of the Aaron Formation and capped by metabasalt of the Virgilina Formation (Glover & Sinha, 1973); (Harris & Glover, 1988). In 1995, Samson, Hibbard, and Wortman published research results indicating that the Virgilina “sequence is composed of Photograph 6-4: MW-5BR (44.5-47.0 ft BGS). Mafic bedrock with mylonitic texture with calcite stringers. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 6-6 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx juvenile, largely mantle-derived crust and likely represents a mature arc built upon an oceanic substrate” (Samson, Hibbard, & Wortman, 1995). At the Mayo Site, fractures were observed within the bedrock at each bedrock core hole. The majority of fractures were relatively small (e.g., close and tight) and appeared to be limited in connectivity between borings. Yields from pumping or packer testing were low. Mylontic textures were observed in some bedrock cores, most notably, in MW-8BR. A zone of highly fractured bedrock was encountered within the competent rock at this location, possibly indicating a local zone of faulting or shear. This zone also yielded small amounts of groundwater. Soil and Rock Mineralogy and Chemistry 6.1.4 Mineralogy and chemistry of the soil and rock encountered are presented in Section 7.0. 6.2 Site Hydrogeology According to LeGrand (2004), the soil/saprolite regolith and the underlying fractured bedrock represent a composite water-table aquifer system. The regolith provides the majority of water storage in the Piedmont province, with porosities that range from 35 percent to 55 percent (Daniel & Dahlen, 2002). Calculated porosities specific to the Site (21 percent to 54 percent) are consistent with this range. Two major factors that influence the behavior of groundwater in the vicinity of the Site include the thickness (or occurrence) of saprolite/regolith and the hydraulic properties of underlying bedrock. Thickness of the regolith is directly related to topography, type of parent rock, and weathering. Topographic highs typically exhibit to thinner soil-saprolite zones, while topographic lows typically exhibit thicker soil-saprolite zones. Saprolite thickness at the Mayo Plant ranges from zero to 66 feet (MW-12 location). LeGrand (2004) makes the generalization that gneiss and schist, which are dominant rock types at the Site, yield thicker soils and moderate to relatively high fracture densities compared with the densities of unaltered igneous rocks such as granite. According to Daniel and Dahlen (2002), foliated rocks such as schist provide planes of weakness that facilitate fracturing at the onset of weathering. This weathering process can produce a relatively transmissive zone. Massive igneous/meta-igneous parent rocks such as granitic gneiss that do not provide tightly spaced planes of weakness and are less susceptible to secondary porosity development due to weathering. Hydrogeologic conditions encountered above these rocks revealed less-distinct transition zones than those at mica schist locations. Porosity of the regolith is directly influenced by parent 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 6-7 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx rock type based on susceptibility to weathering. As weathering advances to formation of clays from mica content, the relative permeability will be reduced. Hydrostratigraphic Layer Development 6.2.1 Hydrostratigraphic units were identified using the framework described by LeGrand (2004) where the soil/saprolite regolith and the underlying fractured bedrock represent a composite water-table aquifer system. Continuous core drilling techniques were employed to continually observe the subsurface for saturated zones, weathered in-situ material, and characteristics of underlying parent rock that may contribute to a water-bearing zone. Borings were advanced to a depth of 50 feet beyond the top of competent bedrock to define water- bearing zones within, adjacent to, and underlying the ash basin. Determination of regolith saturation, transition zone thickness, and potential well yield were made by the field geologist. Four hydrostratigraphic units were identified at the Mayo Site. Those units include the ash pore water (confined to the area of the ash basin); the shallow or surficial zone (alluvium, residual soil, fill/reworked soil, and saprolite); a transition zone between surficial materials and competent bedrock; and fractured bedrock. A description of each is provided in the following section. Hydrostratigraphic Layer Properties 6.2.2 Ash Pore Water The ash pore water unit consists of saturated ash material. Ash depths range from a few feet to approximately 55 feet within the ash harvesting area portion of the ash basin, where approximately 50 feet of ash is saturated. Shallow/Surficial Zone The shallow/surficial flow zone consists of regolith (soil/saprolite) and alluvial material. Thickness of regolith is directly related to topography, type of parent rock, and geologic history. Topographic highs tend to exhibit thinner soil- saprolite zones, while topographic lows typically contain thicker soil-saprolite zones. Alluvium found along Crutchfield Branch was about 7 feet thick and directly overlies saprolite. Saprolite thickness at the Site ranged from not present to more than 50 feet at upgradient well pair MW-12; that thickness of saprolite, however, is an exception beneath the Site. Saprolite beneath the power plant area of the Site and the northern, eastern, and western parts of the Site is almost entirely unsaturated. Saturated saprolite is encountered more frequently in the 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 6-8 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx southern portion of the Site. Alluvium and saprolite are referred to herein as one single unit due to the limited extent of alluvium, the general lack of saturated saprolite, and the interaction of groundwater with surface water. Wells within the shallow flow zone that are installed within alluvial and surficial (shallow) wells contain an “S” designation. Transition Zone The transition zone consists of a relatively transmissive zone of partially weathered bedrock. Observations of core recovered from this zone included rock fragments, unconsolidated material, and highly oxidized bedrock material. Both saturated and unsaturated conditions occur in the transition zone at Mayo. Transition zone wells are labeled with a “D” designation. Fractured Bedrock The fractured bedrock unit occurs within competent bedrock. Bedrock at the Mayo Plant is dominated by granitic gneiss interbedded with phyllite, mica schist and mafic metavolcanic rocks. The majority of water producing fracture zones was found within 50 to 75 feet of the top of competent rock. Water-bearing fractures encountered are only mildly productive (providing water to wells). Bedrock wells are labeled with a “BR” designation. Two transects were selected to illustrate flow path conditions in the vicinity of the ash basin. Section A-A’ is a transverse section through the ash basin, perpendicular to groundwater flow, in relation to the adjacent areas to the northwest and southeast (Figure 6-2). Section B-B’ illustrates conditions upgradient (south) and along the flow path within the ash basin, then downgradient/downstream along the Crutchfield Branch stream valley (Figure 6-3). 6.3 Groundwater Flow Direction As recognized in previous investigations and through compliance monitoring activities, groundwater flow is generally to the north-northeast in the direction of the Crutchfield Branch stream valley. At the Mayo Plant, it is appropriate to combine the flow zones into one generalized flow map. In large portions of the subsurface beneath the Site, the surficial flow zone and the transition zone are not saturated, and the shallow bedrock is the first and only zone where groundwater is encountered. Further, where there are saturated conditions in either regolith or the transition zone, the difference between the water levels in wells in those zones, as compared with the level in adjacent bedrock wells, is miniscule. Horizontal flow dynamics dominate over vertical flow. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 6-9 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx Water levels were measured in Site wells and piezometers within a 24-hour period on June 19-20, 2017 and November 2-3, 2016 to provide water-level elevation data for wet and dry seasons (respectively) at the various flow systems observed at the Site (Table 6- 5). The wet and dry seasons for Mayo were established based on precipitation data from the State Climate Office of North Carolina website (2015) that indicates relatively more precipitation typically occurs during the spring of each year and less precipitation during the fall of each year. Groundwater flow directions and the overall morphology of the potentiometric surface vary little from the “dry” to “wet” seasons. Water levels do fluctuate up and down with significantly increased or decreased precipitation, but the overall groundwater flow directions do not change due to seasonal changes in precipitation. Water level elevation data indicate a north-northeast flow direction in the multiple zones of saturation toward the Crutchfield Branch floodplain (Figures 6-4 through 6- 11). The Mayo Plant ash basin occupies the former stream valley of Crutchfield Branch. The basin acts as an elongated bowl-like feature toward which groundwater flows from the northwest, west, south, and east. A small topographic high is present along the eastern side of the ash basin, and groundwater is somewhat radial away from this feature. Groundwater flow east of the railroad line, which is constructed along a natural ridge, is to the east and toward Mayo Lake. Groundwater flows north-northeast from the ash basin into the small valley formed by Crutchfield Branch. Crutchfield Branch flows north off of the Site into Virginia. Mayo Lake influences groundwater on the east side of the Plant acting as a groundwater discharge area. Groundwater flows from the highest topographic portion of the Site (near the Plant entrance road) to the north and northeast. US Highway 501 follows the crest of a ridge just west of the western edge of the ash basin and influences flow on this portion of the Site. The groundwater flow system at the Site serves both to store and provide a means for groundwater movement. The porosity of the regolith is largely controlled by pore space (primary porosity), whereas in bedrock porosity is largely controlled by the number, size and interconnection of fractures. As a result, the effective porosity in the regolith is normally greater than in the bedrock and thus the quantity of groundwater flow will be greater in the regolith. At the Mayo Site, saturated regolith was observed in only a few wells, and the regolith is the least transmissive of the flow zones. The majority of groundwater across the Site appears to flow through the transition zone and bedrock. Downgradient of the Mayo Plant, groundwater gradients in the shallow flow 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 6-10 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx zone are affected by man-made features (rail cuts, basins, stormwater run-off ditches) and the ash basin. 6.4 Hydraulic Gradient Horizontal hydraulic gradients based on the June 2017 water-level measurements in the various hydrogeologic zones vary across the Site (Table 6-6). Hydraulic gradients, vertical and horizontal, have been calculated along transects that follow cross section B–B’. Horizontal hydraulic gradients were derived by calculating the difference in hydraulic head over the length of the flow path between two wells with similar well construction (e.g., wells within the same water-bearing unit). The following equation was used to calculate horizontal hydraulic gradient: i = dh / dl i = hydraulic gradient; dh = difference between two hydraulic heads (measured in feet); and dl = flow path length between the two wells (measured in feet). Generally horizontal gradients along the southern portion of the Site (B-B’) range from 0.017 feet/feet to 0.018 feet/feet. Horizontal gradients along the northern end of the Site (B-B’) range from 0.03 feet/feet to 0.034 feet/feet. The hydraulic gradient in the northern portion of the Plant is likely due to the much higher relief between this area and the ash basin dam. Vertical hydraulic gradients were calculated by taking the difference in groundwater elevation in a deep and shallow well pair over the difference in total well depth of the deep and shallow well pair. A positive output indicates downward flow and a negative output indicates upward flow. Vertical gradients at select well pairs have been calculated and are presented on Table 6-7 and visually presented in Figure 6-12. Throughout the Site, vertical gradients in saprolite, transition zone and bedrock wells are near equilibrium indicating that there is no distinct horizontal confining layer beneath the Mayo Plant. The approximate range of hydraulic gradient varies from 0.2682 feet/feet to -0.1333 feet/feet. Generally, there is recharge (downward gradients) on the northeast portion of the Site and discharge (upward gradients) to the west in and around the ash basin, with the exception of recharge at the ABMW-2 well cluster. Upward vertical gradients from bedrock, as groundwater from the west, south, and east recharge the groundwater beneath the basin into the former Crutchfield Branch stream valleys, reduces the potential for downward migration of constituents of interest (COI) into bedrock. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 6-11 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx Beyond the ash basin area, the area near the MW-8 cluster has the greatest downward gradient of 0.2689 feet/feet. Relatively strong upward gradients occur near the MW-18 cluster and CW-3/MW-3BR pairs. The greatest upward gradient is at MW-18D and MW-18BR, with a gradient of -0.0627 feet/feet. 6.5 Hydraulic Conductivity Hydraulic conductivity values for the various hydrogeologic zones in which the wells are screened varied Site-wide as determined by the slug test method conducted in accordance with GAP Section 7.1.4 (Table 6-8). Slug test field and analytical methods are discussed in Section 6.8. Hydraulic conductivity values for wells screened in saprolite have a geometric mean of 1.48 x 10-4 cm/sec. Hydraulic conductivity values recorded for wells screened in the transition zone have a geometric mean of 3.31 x 10-4 cm/sec. These measurements reflect the dynamic nature of the transition zone, where hydrologic properties can be heavily influenced by the formation of clays and other weathering by-products. Hydraulic conductivity results for bedrock wells across the Site have a geometric mean of 6.66 x 10-5 cm/sec. The hydraulic conductivity measurements in bedrock wells can be regarded as a generalized representation of the localized bedrock fractures in specific areas of a well cluster. 6.6 Groundwater Velocity To calculate the velocity that water moves through a porous medium, the specific discharge, or Darcy flux, is divided by the effective porosity, ne . The result is the average linear velocity or seepage velocity of groundwater between two points. Groundwater flow velocities for the surficial and transition flow zones were calculated using Darcy's Law equation which describes the flow rate or flux of fluid through a porous media by the following formula: 𝐕𝐬=𝐊𝐢/ ne Vs = seepage groundwater velocity; K = hydraulic conductivity; i = the horizontal gradient; and ne = effective porosity Effective porosities were calculated using laboratory testing and physical soil data presented in Table 6-1 and estimated on a Fetter-Bear diagram (Johnson, 1967). This technique provides a simple method to estimate specific yield; however, there are 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 6-12 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx limitations to this method that may not provide an accurate determination of the specific yield of a single sample (Robson, 1993). Groundwater velocities calculated for the four flow paths described in Section 6.2.2 range from 14.5 feet per year to 25.4 feet per year (Table 6-6). For each flow zone, the geometric mean from the calculated hydraulic conductivity from slug tests was utilized to compute velocity (Appendix C). From background well MW-12S to the ash basin ABMW-3S, there is a horizontal gradient of 0.017 feet/feet with a velocity of 14.5 feet/year; from the same surficial ash basin well (ABMW-3S) to the surficial downgradient well (MW-16S) the horizontal gradient is 0.034 feet/feet and a velocity of 23.3 feet/year. The larger velocity indicated from the ash basin to the Crutchfield Branch floodplain is due to the potentiometric head created by the ash basin being constructed over 100 feet higher than the stream channel into which it flows. Velocities measured across the Plant in the transition zone range from 21.2 feet/year to 25.4 feet/year. At Mayo, groundwater movement in the bedrock flow zone is due primarily to secondary porosity represented by fractures in the bedrock. Primary (matrix) porosity is negligible; therefore, it is not technically appropriate to calculate groundwater velocity using effective porosity values and the method presented above. Bedrock fractures encountered at Mayo tend to be isolated with low interconnectivity. Further, hydraulic conductivity values measure the fractures immediately adjacent to a well screen, not across the distance between two bedrock wells. Groundwater flow in bedrock fractures is anisotropic and difficult to predict, and velocities change as groundwater moves between factures of varying orientations, gradients, pressure, and size. For these reasons, bedrock groundwater velocities calculated using the seepage velocity equation are not representative of actual Site conditions and were not calculated. For additional information on the movement of groundwater around and downgradient of the ash basin over time, refer to discussion concerning groundwater fate and transport modeling (Section 13.0). 6.7 Contaminant Velocity The degree of migration, retardation, and attenuation of constituents in the subsurface is a function of physical and chemical properties of the media through which the groundwater passes. Contaminant velocity depends on factors such as the rate of groundwater flow, the effective porosity of the aquifer material, and the soil-water partitioning coefficient, or Kd term. Soil samples were collected and analyzed for grain size, total porosity, soil sorption (Kd), and anions/cations to provide data necessary for completion of a fate and transport model. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 6-13 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx Constituents enter the ash basin system in both dissolved and solid phases, and those constituents may undergo phase changes that include dissolution, precipitation, adsorption, and desorption. Dissolved phase constituents may undergo these phase changes as they are transported in groundwater flowing through the basin. Phase changes are collectively addressed by specifying a linear soil-groundwater partitioning coefficient (sorption coefficient [Kd]). In the fate and transport model, the entry of constituents into the ash basin is represented by a constant concentration in the saturated zone (pore water) of the basin, and is continually replaced by infiltrating recharge from above. Laboratory Kd terms were developed by University of North Carolina – Charlotte (UNCC) researchers via column testing of 14 site-specific samples of soil. The methods used by UNCC and Kd results obtained from the testing are presented in Appendix C. The Kd data were used as an input parameter to evaluate constituent fate and transport through the subsurface. Boron is relatively mobile in groundwater and is associated with low Kd values. This is primarily because boron is mostly inert, has limited potential for sorption, and lacks an affinity to form complexes with other ions. In general, the low Kd measured for boron allows the constituent to move with groundwater. The higher Kd values measured for other constituents, like arsenic and cobalt, are consistent with the observed, limited migration of these constituents. Constituents like cobalt and arsenic have much higher Kd values and will move at a much slower velocity than groundwater as it sorbs onto surrounding soil. It should be noted that the fractured bedrock flow system is highly heterogeneous in nature and low permeability zones predominate. Geochemical mechanisms controlling the migration of constituents are discussed further in Section 13.0. Groundwater modeling to be performed for the updated CAP will include discussion of contaminant velocities for the modeled constituents. 6.8 Slug Test and Aquifer Test Results As previously discussed, hydraulic conductivity values for the various hydrogeologic zones in which the wells are screened varied Site-wide as determined by the slug test method conducted in accordance with GAP Section 7.1.4 (Table 6-8). Slug test field and analytical methods are included in Appendix G, and results are presented in Appendix C. Slug tests were conducted for all wells installed for the 2015 groundwater assessment except in cases where there was not a sufficient amount of water in the well for the test. Hydraulic conductivity results for the slug tests are summarized below. Where multiple tests were conducted for the same well, the geometric mean result is used. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 6-14 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx Saturated alluvium sediment is scarce throughout the Site and consists of fine to course sand mixed with fine to course gravel. One well, MW-16S, was screened across the saturated alluvium and the hydraulic conductivity is 6.09 x 10-6 cm/sec. Slug test data was analyzed for CAMA and other regulatory wells were screened across the surficial, transition, and bedrock zones throughout the Site. Hydraulic conductivity values for wells screened in saprolite have a geometric mean of 1.48 x 10-4 cm/sec. Hydraulic conductivity values recorded for wells screened in the transition zone have a geometric mean of 3.31 x 10-4 cm/sec. Hydraulic conductivity results for bedrock wells across the Site have a geometric mean of 6.66 x 10-5 cm/sec. Infiltration tests using Guelph permeameters were not performed because the groundwater model developer indicated that those data would not be needed because slug test data were available. Shelby tube samples were collected at nine locations and were used for vertical hydraulic conductivity tests, each conducted on media from five distinct zones: saprolite, residual soil, ash, alluvium, and fill from the toe of the ash basin (Table 6-9). The vertical conductivities were calculated to be, on average, three to four orders of magnitude smaller than the horizontal results. These data indicate that lateral groundwater flow predominates over vertical flow at the Site. 6.9 Fracture Trace Study Results Fracture trace analysis is a remote sensing technique used to identify lineaments on topographic maps and aerial photography that may correlate to locations of bedrock fractures exposed at the earth’s surface. Although fracture trace analysis is a useful tool for identifying potential fracture locations, and by extension, potential preferential pathways for infiltration and flow of groundwater near a site, results are not definitive. Lineaments identified as part of fracture trace analysis may or may not correspond to actual locations of fractures exposed at the surface, and if fractures are present, it cannot be determined from fracture trace analysis whether these are open or healed. Strongly linear features at the earth’s surface are commonly formed by weathering along steeply dipping to vertical fractures in bedrock. Morphological features such as narrow, sharp-crested ridges, narrow linear valleys, linear escarpments, and linear segments of streams otherwise characterized by dendritic patterns are examples. Linear variations in vegetative cover are also sometimes indicative of the presence of exposed fractures, though in many cases these result from unrelated human activity or other geological considerations (e.g., change in lithology). 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 6-15 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx Straight (as opposed to curvilinear) features are commonly associated with the presence of steeply dipping fractures. Curvilinear features in some cases are associated with exposed moderately-dipping fractures, but these also can be a result of preferential weathering along lithologic contacts, or along foliation planes or other geologic structure. As part of this study, only strongly linear features were considered, as these are far more commonly indicative of steeply dipping or vertical fractures. The effectiveness of fracture-trace analysis in the eastern United States, including in the Piedmont, is commonly hampered by the presence of dense vegetative cover, and often extensive land-surface modification owing to present and past human activity. Aerial- photography interpretation is most affected, as identification of small-scale features can be rendered difficult or impossible in developed areas. Available geologic maps for the area were consulted prior to performance of aerial- photography and topographic-map interpretation, to identify lithologies and geologic structure in the area that can control fracture occurrence and orientations. Fracture trace analysis was performed in the vicinity of the Site, and no major faults or shear zones were identified (Figure 6-13). The fracture trace analysis indicates a predominance of southwest-northeast trending features, especially the former Crutchfield Branch stream valley directly below the ash basin. Typical of the Piedmont, joint sets perpendicular to the southwest-northwest trend were also prevalent. Minor north-south trending features were noted. The observations were corroborated by direct field observations and mapping of surface exposures in the area around Mayo Plant and Mayo Lake. Measured (using a Brunton pocket transit) joint set orientations and dominant foliation trends in rocks near the Plant ranged from approximately N30E to N65E. Secondary joint sets were measured with northwest-southeast and north-south trend lines. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 7-1 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx 7.0 SOIL SAMPLING RESULTS Soils samples were collected and tested in accordance with GAP (Section 7.1.1) and the analytical methods for testing soil are summarized in Table 6-3. Soils test data is included in Appendix C. Soil borings were conducted in upgradient and downgradient areas of the ash basin to collect soil samples from the unsaturated zone and the zone of saturation for these areas (Figure 2-8). Physical property testing of soil and saprolite indicate that Site soils are predominantly sand- and silt-sized (sandy loam) with some coarse sandy loams, silty loams, and minor clay inclusions (Table 6-1). Mineralogical analysis of soil samples indicate clay minerals (illite, smectite, and kaolinite) comprise the bulk portion of Site soils, along with quartz, feldspars, and amphiboles (Table 6-2). 7.1 Background Soil Data Four upgradient borings (SB-1, SB-2, SB-3, and SB-4) and groundwater monitoring wells (MW-10BR, MW-11BR, MW-12S, MW-12BR, and MW-14BR) were originally installed for use as background wells and borings. A background soil dataset based on the 2015 CSA data was provided to NCDEQ on May 26, 2017 for consideration of background soil concentrations. Additionally, the revised Statistical Methods for Developing Reference Background Concentrations for Groundwater and Soil at Coal Ash Facilities (statistical methods document) (HDR and SynTerra, 2017) was provided to NCDEQ as a basis for determination. On July 7, 2017, NCDEQ provided a response letter for each Duke Energy coal ash facility that identified soil and groundwater data appropriate for inclusion in the statistical analysis to determine PBTVs for both media. NCDEQ requested that Duke Energy collect a minimum of 10 valid background samples, rather than the previously planned eight that was provided, prior to the determination of PBTVs for each constituent. In addition, soil samples meeting the following criteria are considered valid for use in statistical determinations of PBTVs: Sample was collected from a location that is not impacted by coal combustion residuals or coal associated materials. Sample was collected from a location that is not impacted by other potential anthropogenic sources of constituents. Sample was collected from the unsaturated zone, greater than one foot (ft) above the seasonal high water table elevation. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 7-2 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx NCDEQ determined samples collected from several locations/depths do not meet NCDEQ Inactive Hazardous Site Branch (IHSB) Guidance requirements; therefore, they are not appropriate for use in determining PBTVs. The background soil dataset included laboratory reporting limits for antimony and thallium above the NCDEQ IHSB PSRG Protection of Groundwater values (dated October 2016). NCDEQ requested the values for antimony and thallium be reported below the PSRG Protection of Groundwater values. To address these requirements, additional soil samples were collected from background locations on July 18, 2017 (Figure 2-8). Boring logs associated with the additional soil samples are included in Appendix F. The updated background dataset was screened for outliers prior to statistical determinations. Soil PBTVs were submitted to NCDEQ and partially approved on September 1, 2017 (NCDEQ, September 2017; Appendix A). PBTV values were accepted for all constituents except copper, iron, manganese, sodium, and thallium. PBTVs were recalculated and concurrence on values was achieved in a meeting on October 13, 2017 between NCDEQ, Duke Energy, and SynTerra. PBTVs for soil constituents were computed and are provided in Table 7-1. A background summary report for soils is included as Appendix H. 7.2 Facility Soil Data Soil samples were collected during CSA monitoring well installations. Comparison of soil analytical results with background is discussed below based on the area of the Site. Soil Beneath Ash Basin The contact between the ash and underlying soils in the ash basin borings was visually distinct. There was no visible evidence of substantial migration of ash into underlying soils or mixing of ash with those soils. Five soil samples were collected from three boring/monitoring well locations below the ash basin. Arsenic, beryllium, barium, boron, calcium, chromium, and strontium exceeded the PBTV in at least one soil sample (Appendix B, Table 4). Chromium was the only constituent detected in a concentration that exceeded the Industrial PSRG. Six constituent concentrations (arsenic, chromium, cobalt, iron, manganese, and vanadium) exceeded the POG in the soil below the ash harvesting area (dewatered portion of the ash basin). SPLP was used to determine the ability of simulated rainwater to leach site-specific constituents out of the soil to groundwater. The 2L/IMAC standards are used for reference only of SPLP data; SPLP test results do not represent groundwater; therefore, 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 7-3 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx comparison to 2L/IMAC is not required. The SPLP analyses revealed that only chromium, cobalt, iron, manganese, and vanadium leach from soils beneath the ash basin at concentrations that exceed the 2L/IMAC. Further, those exceedance concentrations were only in one location beneath the ash stack area (AMBW-02BR). Cobalt, iron, manganese, and vanadium appear to be ubiquitous across the Mayo Site in soils regardless of location (e.g., beneath ash, upgradient, downgradient) and tend to leach in concentrations that are often greater than the 2L/IMAC in the leachate even for soils not beneath the ash basin. Soil Beyond Waste Boundary and Within Compliance Boundary Soil samples were collected from soil borings and monitoring wells outside of the ash basin waste boundary (downgradient, sidegradient). Arsenic and chromium exceeded the Industrial PSRG in these soil samples. Detected concentrations of chromium, cobalt, iron, manganese, and vanadium in multiple soil samples exceeded the PSRG POG. Detected concentrations of selenium (three samples) and thallium (four samples) exceeded the PSRG POG in one soil sample each. SPLP results for soils beyond the waste boundary indicate that cobalt, iron, manganese, and vanadium readily leach from natural soils. Leaching of chromium from natural soils is inconsistently observed. Detected soil concentrations of beryllium, boron, cadmium, calcium, chloride, chromium, copper, lead, manganese, mercury, nickel, strontium, thallium, and vanadium exceeded a background soil PBTV in at least one sample. Detections of COIs in soil were sporadic and inconsistent and did not indicate a source of soil impact beyond the ash basin waste boundary. Comparison of PWR and Bedrock Results to Background Three samples were collected from the transition zone or bedrock and analyzed as soil samples. MW-12D (88.5–90) was collected from weathered mica schist. MW-13BR (52- 54) was silt with sand and rock fragments collected within the transition zone. MW- 16BR (54.5-55.5) was dark gray silt collected from within a fracture zone in phyllite. Chromium exceeded the Industrial PSRG and POG for each of three samples. Cobalt, iron, manganese, and vanadium exceeded the PSRG POG in all three samples. Beryllium, cadmium, chromium, and nickel concentrations in MW-12D (88.5-90) exceeded the PBTV. Boron and calcium exceeded the PBTV in MW-16BR (54.5-55.5). Secondary Sources For soil samples beneath the ash basin, only arsenic and strontium concentrations in one soil sample exceeded both the calculated soil PBTV and the NCDEQ Preliminary Soil Remediation Goals (PSRG) Protection of Groundwater (POG) value. Strontium, 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 7-4 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx which does not have a specific PSRG POG, was also detected in the same soil sample at concentrations greater than the PBTV. Arsenic is not detected in groundwater at concentrations greater than the 2L beyond the ash basin waste boundary. Strontium is present above the groundwater PBTVs beyond the compliance boundary in the surficial aquifer only. No other COIs were detected in soil beneath the ash basin at concentrations greater than both a PBTV and POG value. Analysis of soil analytical data presented in Appendix B, Table 4 and Table 7-2 shows that only in a limited extent have COIs from the source mobilized and sorbed onto soils beneath the ash basin. Arsenic and strontium were detected at concentrations greater than their respective PBTV or PSRG, whichever is higher. Figure 7-1 shows the exceeding data in relation to the ash basin. Although there are limited exceedances of PBTVs and/or PSRGs in soil beneath the ash basin, the distribution of COIs does not appear widespread beneath the ash basin area. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 8-1 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx 8.0 SEDIMENT RESULTS Sediment samples were collected during the CSA from 10 locations beyond the perimeter of the ash basin (Figure 2-8). Sediment analytical results are presented in Appendix B, Table 5. 8.1 Sediment/Surface Soil Associated with AOWs Seven sediment sample locations were co-located with designated Areas of Wetness (AOWs). For some of these locations, solid material was collected at or near the point of emergence or flow of water. In most cases, the solid material that was collected was actually surface soil over which water originating at the AOW was flowing. For sample locations S-3 and SW-CB1, sediment was collected from the bottom of the channel. A description of the AOW and the results of sediment analysis are provided below: S-1: Engineered toe drain system (west toe drain) draining into a concrete structure. Sediment was collected from the bottom of the structure. Arsenic, chromium, iron, and manganese concentrations exceeded the PSRG POG. Boron was also detected in this sediment sample. Arsenic, boron, iron, and zinc concentrations exceeded the respective soil PBTV. S-2: Engineered toe drain system (east toe drain) draining into a concrete structure. Sediment was collected from the bottom of the structure. Chromium, cobalt, iron, manganese, and vanadium concentrations exceeded the POG. There were no concentrations that exceeded a PBTV. S-2B: Adjacent to the S-2 structure in an area of rip rap at the toe of the east side groin of the ash basin dam. Flow appears to originate several hundred feet upslope within the groin. Sediment was collected from beneath/around rip rap. There were no concentrations that exceeded a PBTV. Cobalt, iron, and manganese concentrations exceeded the PSRG POG. S-3: Located approximately halfway between the east toe drain (S-2) and Mayo Lake Road in the former Crutchfield Branch valley and downstream of the confluence of water originating from S-1 area, S-2 area, and S-8. Chromium, cobalt, iron, manganese, and vanadium concentrations exceeded the PSRG POG only. Chromium and lead concentrations exceeded the respective PBTVs. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 8-2 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx S-4: Adjacent to S-1 in a wide, shallow channel comprised mostly of flow from S-1. Water was typically stagnant with abundant iron oxide floc and algae growth. Sediment collected from channel. Arsenic, chromium, cobalt, iron, manganese, and vanadium concentrations exceeded both the PSRG POG. Boron was also detected in this sediment sample. Arsenic, boron, iron, and manganese concentrations exceeded the respective PBTVs. S-8: Located in hillside adjacent to ash basin dam. Flow emerges from hillside at several points, merges with diffuse flow along hillside into a single “channel,” and flows downhill. Sediment collected from inundated surface soil at the point of water emergence. Chromium, cobalt, iron, manganese and vanadium exceeded the POG but only manganese exceeded a PBTV. SW-CB1: Located in the former Crutchfield Branch valley downstream of S-3 at the culvert under Mayo Lake Road. Chromium, cobalt, iron, manganese, and vanadium concentrations exceeded the POG; however, when compared to soil PBTVs, none of the detected concentrations exceeded the respective PBTVs. The detected concentrations of shallow, sediment/inundated surface soils that are laterally restricted in areal extent do not indicate a source of impact to groundwater. 8.2 Sediment in Major Water Bodies Three sediment sample locations were located in surface water bodies. For these locations, sediment was collected from the bottom of the stream channel. As described below, two of the locations are upstream reference locations and one of the locations is downstream of the ash basin. A description of the sediment sampling location and the results of sediment analysis are provided below: SW-REF1: Reference location in wooded area south of the power plant and upstream from the ash basin. This location is downstream of NPDES permitted stormwater outfall 010. This location is upstream from S-6. The stream is approximately 3 feet wide with a silt and cobble bottom substrate. Chromium, iron, manganese, and vanadium concentrations exceeded the POG; however, when compared to soil PBTVs, none of the detected concentrations exceeded the respective PBTVs. S-6: Reference location in wooded area south of the power plant and upstream from the ash basin. This location is downstream of NPDES permitted stormwater outfall 010. The stream is approximately 6 feet wide with a silt, floc, and cobble 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 8-3 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx bottom substrate. Chromium, cobalt, iron, manganese, and vanadium concentrations exceeded the POG; however, when compared to soil PBTVs, none of the detected concentrations exceeded the respective PBTVs. SW-CB2: Located in Crutchfield Branch north of the Mayo Plant on privately owned property. Chromium, cobalt, iron, manganese, and vanadium concentrations exceeded the POG; however, when compared to soil PBTVs, none of the detected concentrations exceeded the respective PBTVs. The detected concentrations of COIs in sediment from water bodies do not indicate a source of impact to groundwater. In fact, exceedances of PSRGs in upstream sediment samples mirror those of samples taken downstream of the ash basin. None of the detected constituents, with the exception of chromium in S-3, exceeded a PBTV for soil. Detected concentrations for sediment from reference locations are similar, and in some cases higher than, those from downstream locations. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 9-1 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx 9.0 SURFACE WATER RESULTS The Mayo Plant ash basin is located in the central part of the Site. The ash basin receives surface water runoff and groundwater recharge from upland areas that are northwest, west, south, and southeast of the ash basin. Groundwater from the ash basin discharges downgradient of the ash basin dam into the former Crutchfield Branch valley. Flow from the west toe drain (S-1) follows a small channel to the north (S-4) and then east where it merges with the channel from the east toe drain (S-2). Several AOWs, in addition to the toe drains (S-1 and S-2), emerge beneath or near the ash basin dam (S- 1A, S-2A, S-2B, S-8, and S-10). Flow from AOWs S-2A and S-2B merge with the flow from the east toe drain in the channel downstream of the toe drain. Flow from S-8 merges with the flow originating from the east toe drain prior to it merging with the flow originating at the west toe drain. The combined flow re-forms (represented by sample S-3) and flows north toward and beneath Mayo Lake Road (sample location SW- CB1). S-10 has only been noted to be flowing one time and has been dry during each subsequent observation/sampling attempt. Flow from S-10 would enter the system just upstream from AOW location SW-CB1. Crutchfield Branch reforms north of the Plant property boundary and flows through a heavily beaver-impounded channel toward the Virginia state line (surface water sample SW-CB2). A small intermittent tributary to Crutchfield Branch forms north of the 1981 C&D Landfill and flows north beneath Mayo Lake Road (SW-CBT1), eventually merging with Crutchfield Branch off-site. Aqueous samples discussed within the following sections include three distinct types: 1) ash basin wastewater, 2) AOWs, and 3) named surface waters. For the scope of this CSA, it is only appropriate to compare named surface waters to NCDENR Title 15A, Subchapter 02B Surface Water and Wetland Standards (2B) because AOWs, wastewater and wastewater conveyances (effluent channels) are evaluated and governed wholly separate in accordance with the NPDES Program administered by NCDEQ DWR. This process is on-going in a parallel effort to the CSA and subject to change. Surface water and AOW analytical results are included in Appendix B, Table 2, and Table 3. The surface water sample locations are included on Figure 2-8. Ash Basin Water Samples Sample S-5 was collected directly from the water column in the ash basin adjacent to the wastewater treatment area and prior to water flowing into the forebay. Water in the ash basin is a combination of groundwater that may have discharged into the basin, water 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 9-2 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx column that is in contact with ash present in the bottom of the ash basin, local storm water runoff, cooling tower blowdown, water sourced from permitted internal outfalls, and treated wastewater. Ash sluicing was discontinued in October 2016. Area of Wetness (AOW) Sample Locations Eleven AOWs have been identified and sampled routinely for monitoring purposes. The Mayo Site is inspected semi-annually for the presence of existing and potentially new AOWs along the former Crutchfield Branch valley and along and downgradient of the ash basin. Inspections include observations of the ash basin along the toe of the dam; areas below full pond elevation for the ash basin; between the ash basin and receiving waters; and drainage features associated with the basin including engineered channels. Per the interim administrative agreement, these inspections are governed by a Discharge Identification Plan (DIP) until the NPDES permit is issued. These AOWs include: Two engineered toe drains (S-1 and S-2) at the base of the ash basin dam, Four AOW locations in/around the ash basin (S-1A, S-2B, S-8, and S-10), Three AOW locations within the channels formed from the combined flow of the toe drain outfalls (S-3, S-4, SW-CB1), One location from an intermittent flow in a flow regime distinct from the ash basin (SW-CBT1), and One location upstream/upgradient of the ash basin (S-9). Surface Water Sample Locations Three surface water samples were also collected and analyzed. SW-CB2 was collected from Crutchfield Branch on private property north of the Plant property boundary and at the North Carolina/Virgina state line. SW-REF1 and S-6 were collected from the southern portion of the Mayo Site, upstream and upgradient from the ash basin, from a small stream that originates south of the coal pile area, flows beneath the Plant entrance road, and eventually to Mayo Lake. This stream channel is associated with NPDES stormwater outfall 010. SW-REF1 is located about 100 feet upstream of where a previously identified, but now seasonally dry, AOW (S-7) merged with the small stream. S-6 is situated several hundred feet upstream from the point that the stream enters Mayo Lake. These two locations were intended for use as reference locations for other Site surface water locations. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 9-3 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx NCDEQ Sample Locations NCDEQ directed the sampling and analysis of AOWs in March 2014. The locations and analytical results from this sampling event were provided by NCDEQ to Duke Energy and are assumed to be accurate. Assessment samples collected during the CSA were analyzed for a more exhaustive constituent list than those collected in 2014 by NCDEQ and Duke Energy (for example, cobalt was not analyzed during the 2014 sampling efforts). 9.1 Discussion of Results for Constituents Without Established 2B Standards A 2B has not been established for a number of constituents. A summary of these results for COIs without 2B standards follows. The results from surface water location SW-CB2 are compared to the upgradient, reference surface water data from locations SW-6 and SW-REF1. The background surface water concentrations have not been statistically derived or approved by NCDEQ and are for discussion purposes only. Boron is detected in SW-CB2 but not in reference locations. Strontium was detected in both SW-CB2 and in reference location SW-REF1; however, the detected concentrations in SW-CB2 were generally two to three times higher than in SW-REF1. Other constituents are higher in SW-CB2 surface water than in the reference locations (e.g., TDS, cobalt, iron, manganese); although, not several times higher. Chromium was only detected in a reference location (SW-REF1). Aluminum and vanadium concentrations were higher in the reference locations than in the downstream surface water location (SW-CB2). 9.2 Comparison of Exceedances of 2B Standards The following surface water location occurs in Crutchfield Branch and sample results are compared to 2B (Class B) values. SW-CB2 Surface water sample results from upstream, reference locations that eventually flow to Mayo Lake are compared to 2B (Class WS-V) values. S-6, SW-REF1 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 9-4 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx A review of historic surface water sample data indicates that only turbidity and dissolved oxygen (DO) concentrations have exceeded a 2B standard. Specifically, the historic data indicate the following 2B exceedances: S-6: DO (1 of 8 results); Turbidity (1 of 8 results) SW-REF1: DO (3 of 6 results) SW-CB2: • DO (3 of 6 results) • Dissolved copper was reported above the 2B in one sample collected/ analyzed in April 2017. 9.3 Discussion of Surface Water Results As previously described, prior to construction of the ash basin, Crutchfield Branch was a perennial stream that originated about 1,000 feet southwest of the current ash basin footprint. The ash basin now encapsulates the former headwaters of Crutchfield Branch and two smaller, intermittent streams that flowed into Crutchfield Branch. Groundwater underlying the ash basin flows north-northeast and eventually through the base of the dam. Several AOWs and discharges from engineered toe drains emerge from the area around the ash basin dam and contribute to the volume of water flowing downstream of the dam in the former Crutchfield Branch stream valley. The reformed Crutchfield Branch flows northeast from Mayo Lake Road toward the state line. Crutchfield Branch, its valley, and its tributaries are groundwater discharge zones downgradient of the ash basin. The current AOW and surface water data reflect that the majority of the flow in Crutchfield Branch is associated with engineered drainage from the ash basin immediately below the dam and other natural seepage of ash basin water. The groundwater in the area in deeper flow zones near the receiving stream generally contains constituent concentrations less than those of the receiving waters. Boron concentrations are greatest proximate to the engineered toe drains. Boron concentrations in Crutchfield Branch decrease from the Mayo Lake Road to the state line. Manganese concentrations are similarly consistent immediately downstream of the ash basin and begin to decline downstream; however, at the most downstream surface water location (SW-CB2), manganese concentrations increase. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 9-5 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx Piper diagrams, graphical representations of major water chemistry using two ternary plots and a diamond plot, for AOWs and surface water are included as Figure 9-1. One of the ternary plots shows the relative percentage of major cations in individual water samples, and the other shows the relative percentage of the major anions. The geochemical nature of AOW and surface water samples projected on the diagrams coincides closely with monitoring wells MW-3 and MW-16S, both installed within the alluvium of Crutchfield Branch. The geochemical signatures of surface water from upgradient, reference locations also coincide closely with each other and are distinct from AOWs/surface water downstream of the ash basin. Piper diagrams for groundwater monitoring wells are presented and discussed in more detail in Section 10.0. Additional surface water sampling will be performed and an evaluation of potential impacts of groundwater on surface water will be presented in the CAP. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 10-1 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx 10.0 GROUNDWATER SAMPLING RESULTS This section provides a summary of groundwater analytical results for the most recent monitoring event (2Q2017; March-April 2017) and discussion of historical data results and trends. A comprehensive table with all media analytical sample results is provided in Appendix B. As indicated on the comprehensive data table, at the request of NCDEQ, the groundwater results have been marked to indicate data points excluded based on a measured turbidity greater than 10 NTUs; high pH values that may indicate possible grout intrusion into the well screen; and data that may be auto-correlated because it was collected within 60 days of a previous sampling event. The most recent (March-April 2017) valid data collected is presented on the pertinent maps. Two limited rounds of sampling and analysis were conducted in March 2015 and June 2015. They were included in the initial CSA report (SynTerra, 2015a). In addition, the following monitoring events have been completed: Comprehensive Round – September 2015 (reported in CAP, Part 1) Comprehensive Round – December 2015 (reported in CAP, Part 2) Comprehensive Round – January 2016 (reported in CSA Supplement 1) Limited Round (background wells; wells located along flow transect) – April 2016 (reported in CSA Supplement 1) Limited Round (background wells; wells located along flow transect) – July 2016 Limited Round (background wells; wells located along flow transect) – September 2016 Comprehensive Round – November 2016 Comprehensive Round – February 2017 Comprehensive Round – April 2017 Groundwater sampling methods were in general accordance with the procedures described in the GAP (SynTerra, 2014c) and included in Appendix G. Analytical data reports are included in Appendix I. A background summary report for groundwater is included as Appendix H. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 10-2 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx 10.1 Background Groundwater Concentrations Locations for background monitoring wells installed in 2015 for the initial CSA field effort were chosen based on the information available including the previously installed NPDES monitoring well network, horizontal distance from the waste boundary, the relative topographic and groundwater elevation difference compared to the ash basin surface water. After the background wells were installed and a sufficient number of samples were collected, statistical analysis was used to confirm the analytical results represented background conditions. The following monitoring wells have been approved by NCDEQ as background monitoring wells (Zimmerman to Draovitch, July 7, 2017; Appendix A). Background monitoring wells are depicted on Figure 2-8. MW-12S – Surficial BG-2 – Transition Zone MW-12D – Transition Zone BG-1 – Bedrock MW-13BR – Bedrock MW-14BR - Bedrock Monitoring wells BG-1 (transition zone) and BG-2 (bedrock) were installed prior to 2015. Samples have been collected from these wells since 2010, and both wells are currently used as background wells for NPDES and other monitoring programs. Both wells are hydraulically upgradient of the ash basin. Monitoring well pair MW-12S and MW-12D is located approximately 2,500 feet upgradient of the ash basin. Geologic conditions encountered at the MW-12 well pair were different than conditions encountered elsewhere at the Site. Saprolite encountered at MW-12 (approximately 70 feet) was the thickest observed at the Site and the saprolite graded into highly weathered mica schist before grading into weathered gneiss. This portion of the Site was not subjected to significant grading and reworking, and the “natural” thickness of saprolite is still intact. Monitoring well MW-13BR is upgradient of the ash basin on the west side of US Highway 501. MW-13BR was installed on the edge of a topographic draw that extends from the westernmost arm of the ash basin, only about 750 feet to the east. The well was to be used to investigate the potential of a fracture zone presenting itself as a linear topographic draw. However, based on groundwater elevation data, it is apparent that 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 10-3 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx MW-13BR is upgradient of influence from the ash basin and serves as a background bedrock monitoring location. MW-14BR is located on the northwest portion of the Site east of US Highway 501. The hydraulic gradient from MW-14BR to the ash basin is downward approximately 25 feet across the 750 feet of horizontal elevation change. Monitoring wells MW-5BR, MW-10BR, and MW-11BR were originally proposed as Site background wells but have been removed from use as background well. MW-11BR is screened in a rock type that was only encountered at this drilling location, namely a hard, dark green hornblende, chlorite-bearing gabbro. There has been a steady increase in pH since installation in 2015; therefore, the well screen is thought to have been impacted by grout during the well installation process. Bedrock samples from the well screen zone were analyzed for paste pH which indicated a pH of 9.30 for the bedrock. This indicates that the pH values of groundwater measured in the field were likely a result of grout incursion. Therefore, the analytical data for this well is no longer used; however, the well is retained for use as a piezometer only. Background Dataset Statistical Analysis 10.1.1 The revised background groundwater datasets and statistically determined PBTVs are presented below. The current background monitoring well network consists of wells installed within three flow zones – surficial, transition zone, and fractured bedrock. Well locations are presented on Figure 2-8. For groundwater datasets with less than 10 valid samples available for determination of PBTVs, no formal upper tolerance limit (UTL) statistics were run and the PBTV for a constituent and groundwater flow system were computed to be either: The highest value, or If the highest value is above an order of magnitude greater than the geometric mean of all values, then the highest value should be considered an outlier and removed from further use and the PBTV is computed to be the second highest value. This procedure applies to the dataset for the surficial flow zone. NCDEQ requested that the updated background groundwater dataset exclude data from the background data set due to one or more of the following conditions: Sample pH is greater than or equal to 8.5 standard units unless the regional NCDEQ office has determined an alternate background threshold pH for the Site. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 10-4 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx Sample turbidity is greater than or equal to 10 NTUs. Result is a statistical outlier identified for background sample data collected through second quarter 2017. Sample collection occurred less than a minimum 60 days between sampling events. Non-detected results are greater than 2L/IMAC. Statistical determinations of PBTVs were performed in accordance with the revised Statistical Methods for Developing Reference Background Concentrations for Groundwater and Soil at Coal Ash Facilities (statistical methods document) (HDR and SynTerra, 2017). Background datasets provided to NCDEQ on May 26, 2017 were revised based on input from NCDEQ in the July 7, 2017 correspondence. The revised background datasets for each flow system used to statistically determine naturally occurring concentrations of inorganic constituents in groundwater are provided in Table 10-1. The following sections summarize the refined background datasets along with the results of the statistical evaluations for determining PBTVs. Shallow/Surficial Flow Unit One well, MW-12S, monitors background groundwater quality within the surficial (saprolite) flow zone. Eleven samples have been collected from MW-12S since June 2015. Samples collected less than 60 days between sampling events, statistical outliers, and invalid data have been omitted from use in statistical determinations which has resulted in six valid samples for the surficial flow zone. Additional samples will be collected from MW-12S to achieve the minimum requirement of 10 valid samples for determining PBTVs. Only iron, cobalt, manganese, and vanadium currently have a PBTV greater than the 2L/IMAC. Transition Zone Flow Unit Two wells, BG-2 and MW-12D, monitor background groundwater quality within the transition zone. The transition zone background groundwater dataset meets the minimum requirement of 10 samples for all constituents except radionuclides (radium-226, radium-228, uranium-233, uranium-234, uranium-236, and uranium-238). PBTVs for radionuclides were computed to be either the 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 10-5 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx maximum value, or, if the maximum value was above an order of magnitude greater than the geometric mean of all values, the second highest value. PBTVs were calculated for all remaining constituents monitored within the transition zone using formal UTL statistics. Only iron, manganese, and vanadium currently have a PBTV greater than the 2L/IMAC for the transition zone unit. Fractured Bedrock Flow Unit Three wells (BG-1, MW-13BR, and MW-14BR) monitor background groundwater quality within the fractured bedrock. NCDEQ identified four samples collected from BG-1 and two samples collected from MW-13BR that were collected less than 60 days between sampling events and are not appropriate for use in statistical determinations. The background bedrock dataset satisfies the minimum requirement of 10 samples to use formal UTL statistics to derive PBTVs for each constituent monitored within fractured bedrock. Only iron, cobalt, manganese, and vanadium currently have a PBTV greater than the 2L/IMAC for the bedrock flow unit. Summary The calculated groundwater PBTVs were less than their applicable 2L/IMAC for every constituent within each of the three flow units except: Cobalt: PBTVs of 1.02 µg/L (surficial); 1.19 µg/L (bedrock) versus IMAC of 1 µg/L. Iron: PBTVs of 385 µg/L (surficial); 1,319 µg/L (transition zone); 2,550 µg/L (bedrock) versus 2L of 300 µg/L. Manganese: PBTVs of 253 µg/L (surficial); 298 µg/L (transition zone); 544 µg/L (bedrock) versus 2L of 50 µg/L. Vanadium: PBTVs of 0.974 µg/L (surficial); 5.88 µg/L (transition zone); 5.52 µg/L (bedrock) versus IMAC of 0.3 µg/L. Groundwater PBTVs were calculated for the following constituents that do not have a 2L standard, IMAC or Federal Maximum Contaminant Level (MCL) established: alkalinity, bicarbonate, calcium, carbonate, magnesium, methane, potassium, sodium, sulfide, and TOC. Background threshold values will continue to be evaluated and adjusted over time as additional background data becomes available. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 10-6 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx Piper Diagrams (Comparison to Background) 10.1.2 A Piper diagram is a graphical representation of major water chemistry using two ternary plots and a diamond plot. One of the ternary plots shows the relative percentage of major cations in individual water samples and the other shows the relative percentage of the major anions. The apices of the cation plot are calcium, magnesium, and sodium plus potassium. The apices of the anion plot are sulfate, chloride, and carbonates. The two ternary plots are projected onto the diamond plot to represent the major ion chemistry of a water sample. The ion composition can be used to classify groundwater of particular character and chemistry into sub-groups known as groundwater facies. Percentages of major anions and cations are based on concentrations expressed in meq/L (EPRI, 2006). Plots of pore water, surficial, transition, and bedrock groundwater including background locations are shown on Figure 10-1, Figure 10-2, and Figure 10-3, respectively. Historical data from neighboring Orange County serves as a useful comparison to those wells identified as background in this CSA (Cunningham & Daniel, 2001). In Orange County, the groundwater is characterized as circumneutral Ca- HCO3. At Mayo, background wells in the surficial zone and transition zone fall within the same region as those studied in Cunningham and Daniel, supporting their selection as background. In the bedrock flow zone at Mayo, three background wells show Ca-HCO3 waters, while MW-14BR falls outside of the range given by Cunningham and Daniel and is dominated by Na-HCO3. 10.2 Downgradient Groundwater Concentrations The following is a summary of groundwater analytical data for areas around the Mayo ash basin. The comprehensive groundwater analytical data table is included as Appendix B, Table 1. Monitoring Wells Beneath Ash Basin 10.2.1 Monitoring wells ABMW-2BR, ABMW-2BRL, ABMW-3S, ABMW-4D, and ABMW-4BR were installed beneath the ash basin. Since monitoring began in 2015, these wells have one or more detected concentrations greater than PBTVs for the following COIs: pH: ABMW-2BR (PBTV); ABMW-2BRL (PBTV); ABMW-3S (2L); ABMW- 4BR (PBTV); ABMW-4D (PBTV/2L) Arsenic: ABMW-2BR (PBTV); ABMW-2BRL (PBTV); ABMW-3S (PBTV); ABMW-4BR (PBTV); ABMW-4D (PBTV/2L) 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 10-7 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx Barium: ABMW-2BR (PBTV); ABMW-3S (PBTV); ABMW-4BR (PBTV); ABMW-4D (PBTV/2L) Boron: ABMW-3S (PBTV/2L); ABMW-4D (PBTV/2L) Cobalt: ABMW-3S (PBTV/ IMAC); ABMW-4D (PBTV/ IMAC) Hexavalent Chromium: ABMW-2BR (PBTV); ABMW-2BRL (PBTV); ABMW-3S (PBTV); ABMW-4BR (PBTV); ABMW-4D (PBTV) Iron: ABMW-2BR (2L); ABMW-2BRL (2L); ABMW-3S (PBTV/2L); ABMW- 4BR (2L); ABMW-4D (PBTV/2L) Manganese: ABMW-2BR (2L); ABMW-2BRL (2L); ABMW-3S (PBTV/2L); ABMW-4BR (2L); ABMW-4D (PBTV/2L) Molybdenum: ABMW-4D (PBTV) Strontium: ABMW-2BR (PBTV); ABMW-2BRL (PBTV); ABMW-3S (PBTV); ABMW-4BR (PBTV); ABMW-4D (PBTV) Sulfate: ABMW-4BR (PBTV) TDS: ABMW-2BR (PBTV); ABMW-3S (PBTV); ABMW-4BR (PBTV/2L); ABMW-4D (PBTV) Vanadium: ABMW-2BR (IMAC); ABMW-2BRL (IMAC); ABMW-3S (IMAC); ABMW-4BR (IMAC); ABMW-4D (PBTV/ IMAC) COI concentrations in the bedrock wells beneath the ash basin (ABMW-2BR, ABMW-2BRL, and ABMW-4BR) have lower concentrations than wells completed in the surficial and transition zone flow units under the ash basin. Since monitoring began in 2015, the bedrocks wells have had periodic 2L/IMAC exceedances for iron, manganese, TDS, thallium, and vanadium and one or more concentrations greater than PBTVs for pH, arsenic, barium, hexavalent chromium, strontium, sulfate, and TDS. Since their first sampling event in 2015, the wells completed in the saprolite or transition zone (ABMW-3S and ABMW- 4D) have had periodic exceedances above 2L/IMAC for pH, antimony, arsenic, barium, boron, cobalt, iron, manganese, TDS, thallium, and vanadium. In the March-April 2017 sampling event, 2L/IMAC exceedances, for all wells beneath the ash basin, were pH, arsenic, barium, boron, cobalt, iron, manganese, TDS, and vanadium. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 10-8 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx Monitoring Wells Downgradient of Ash Basin 10.2.2 Monitoring wells CW-2, CW-2D, MW-3, MW-16S, MW-16D, and MW-16BR are located directly downgradient of the ash basin in or near the Crutchfield Branch valley. These wells have one or more detected concentrations greater than PBTVs and/or 2L/IMAC as listed below: pH: CW-2 (2L); CW-2D (2L); MW-3 (2L); MW-16S (2L); MW-16D (PBTV); MW-16BR (PBTV/2L) Arsenic: MW-16BR (PBTV) Barium: CW-2 (PBTV); MW-3 (PBTV); MW-16S (PBTV) Boron: CW-2 (PBTV/2L); CW-2D (PBTV); MW-3 (PBTV/2L); MW-16S (PBTV) Chromium: MW-16S (PBTV) Hexavalent Chromium: MW-3 (PBTV) Cobalt: CW-2D (IMAC); MW-16S (PBTV/IMAC) Iron: MW-16S (PBTV/2L); MW-16D (2L); MW-16BR (2L) Manganese: CW-2 (PBTV/2L); CW-2D (PBTV/2L); MW-3 (PBTV/2L); MW- 16S (PBTV/2L); MW-16D (2L); MW-16BR (2L) Molybdenum: MW-3 (PBTV); MW-16D (PBTV); MW-16BR (PBTV) Strontium: CW-2D (PBTV); MW-3 (PBTV); MW-16S (PBTV) Sulfate: CW-2 (PBTV); CW-2D (PBTV); MW-3 (PBTV); MW-16S (PBTV); MW-16D (PBTV) TDS: MW-3 (PBTV); MW-16S (PBTV) Vanadium: CW-2 (IMAC); CW-2D (IMAC); MW-3 (IMAC); MW-16S (IMAC); MW-16D (IMAC); MW-16BR (IMAC) In the March-April 2017 sampling event, 2L/IMAC exceedances for the downgradient monitoring wells in or near the Crutchfield Branch valley were noted only for pH, boron, iron, manganese, and vanadium. Monitoring wells CW-3, CW-4, CW-6, MW-2, and MW-3BR are also located downgradient of the ash basin, although outside of the Crutchfield Branch stream valley. MW-6BR was categorized as a piezometer (water level only) after development proved unsuccessful due to low groundwater yield. These downgradient wells have one or more historic concentrations greater than PBTVs 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 10-9 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx for pH, chromium, cobalt, iron, manganese, sulfate, and TDS. For the March- April 2017 sampling event, 2L/IMAC exceedances were noted for pH, iron, manganese, TDS, and vanadium. Monitoring Wells in Separate Flow Regime 10.2.3 Monitoring wells MW-7D, MW-7BR, MW-8BR, MW-9BR, and MW-9BRL are located east of the ash basin between the ash basin and Mayo Lake. Recently reviewed groundwater elevation data from construction-related piezometer installations for the future FGD Pond confirm these wells to be in a separate flow regime from the ash basin, as the ridge under the railroad serves as a groundwater divide. MW-9BR continues to have turbidity greater than 10 NTUs for every sampling event, resulting in excluded analytical data. As a result, NCDEQ requested a replacement well deeper into bedrock, and the conversion of MW-9BR into a piezometer. MW-9BRL was installed on February 2, 2017. Due to development time prior to sampling, groundwater analytical data was not available for MW-9BRL at the time of this report. MW-7D and MW-7BR are in the footprint of the future FGD Pond requiring abandonment prior to construction activities. The well pair was sampled a total of five times, with the last sampling event on November 4, 2016. NCDEQ DWR was consulted and approved abandonment of the well pair. The wells were abandoned on June 12, 2017. MW-7D, MW-7BR, and MW-8BR have one or more concentrations greater than PBTVs for pH, barium, hexavalent chromium and total chromium, manganese, molybdenum, strontium, sulfate, and TDS. In the March-April sampling event, MW-8BR was the only well sampled in this area, and 2L/IMAC exceedances were pH, iron, manganese, vanadium. Monitoring Wells East of Rail Line (Separate Flow 10.2.4 Regime) Monitoring wells MW-18D, MW-18BR, MW-19D, and MW-19BR were installed northeast of the Plant, between the existing FGD Pond and Mayo Lake, to assess the east side of the railroad line, east of the existing FGD ponds. A groundwater divide roughly bisects the Site property from southwest to northeast. In general terms, the divide appears to be positioned roughly beneath the railroad line. Groundwater flow further to the east has been demonstrated to flow east of the divide into the Mayo Lake flow regime and flow to the west is toward the former C&D landfill area, the tributary on the east side of the landfill, and the ash basin. A monitoring well cluster (MW-18D/BR) was installed directly across the railroad tracks from the FGD ponds (east) between the rail line and the stream that receives NPDES outfall 002 flow from the ash basin. A second well cluster 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 10-10 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx (MW-19D/BR) was installed further to the east and closer to Mayo Lake. CW-1 and CW-1D are compliance wells, in the same area, across the NPDES 002 outfall. These wells have one or more concentrations greater than PBTVs and/or 2L/IMAC for the following COIs: pH: CW-1 (PBTV/2L); MW-18D (2L) Barium: MW-18BR (PBTV) Hexavalent Chromium: MW-18BR (PBTV); MW-19BR (PBTV) Chromium: CW-1D (PBTV/2L) Iron: MW-19D (PBTV/2L); MW-19BR (2L) Manganese: CW-1D (2L); MW-18BR (PBTV/2L); MW-19D (PBTV/2L); MW-19BR (PBTV/2L) Molybdenum: CW-1D (PBTV) Strontium: MW-18BR (PBTV); MW-19D (PBTV); MW-19BR (PBTV) Sulfate: CW-1 (PBTV); MW-18D (PBTV); MW-19D (PBTV); MW-19BR (PBTV) TDS: MW-18BR (PBTV); MW-19D (PBTV/2L); MW-19BR (PBTV/2L) Vanadium: CW-1 (IMAC); CW-1D (IMAC); MW-18D (IMAC); MW-18BR (IMAC); MW-19D (IMAC); MW-19BR (IMAC) Piper Diagrams (Comparison to Downgradient/ 10.2.5 Separate Flow Regime) The Piper Diagrams (Figures 10-1 to 10-3) display water chemistry for pore water and downgradient wells. Groundwater samples from downgradient wells generally indicate calcium-bicarbonate type water, typical of shallow fresh groundwater. A few exceptions to this characterization include CW-2, with a geochemical signature which is representative of calcium-sulfate type water, and background well MW-14BR, with a signature representative of sodium- bicarbonate type water indicative of deep groundwater. As previously described in Section 3.3, pore water at the Mayo Site is atypical when compared to other ash leachate sources. As such, comparison of monitoring well data within the Piper diagram is limited in its usefulness as similar characteristics are also observed in Mayo Site pore water, unimpacted groundwater, and potentially impacted groundwater. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 10-11 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx 10.3 Site-Specific Exceedances (Groundwater COIs) Site-specific COIs were developed by evaluating groundwater sampling results with respect to PBTVs, applicable regulatory standards, and additional regulatory input/requirements. The approach to determining those constituents which should be considered constituents of interest (COI)s for the purpose of this assessment is discussed in the following section. Provisional Background Threshold Values (PBTVs) 10.3.1 As presented in 2L .0202 (b)(3) — “Where naturally occurring substances exceed the established standard, the standard shall be the naturally occurring concentration as determined by the Director” — the following report was provided to NCDEQ: Statistical Methods for Developing Reference Background Concentrations for Groundwater and Soil at Coal Ash Facilities (HDR and SynTerra, 2017). NCDEQ (July 7, 2017) addressed each Duke Energy coal ash facility and identified soil and groundwater data appropriate for inclusion in the statistical analysis to determine BTVs and PBTVs. A revised and updated technical memorandum that summarized revised background groundwater datasets and statistically determined PBTVs for the Mayo Plant was submitted to NCDEQ on August 16, 2017. A list of NCDEQ-approved groundwater PBTVs were provided to Duke Energy on September 1, 2017 (Zimmerman to Draovitch; Appendix A). Each of the proposed groundwater PBTVs, as proposed, was accepted. Applicable Standards 10.3.2 As part of CSA activities at the Site, sampling and analysis for inorganic constituents has been conducted for coal ash, ponded water in the ash basin, ash pore water, AOW, surface water, sediment, soil, and groundwater downgradient/sidegradient of the ash basin and in background areas. Based on comparison of those sampling results from the multiple media to background values and applicable regulatory values, potential lists of COIs were developed in the 2015 CSA, CAPs, and CSA Supplement. For the purpose of developing the groundwater COIs, constituent exceedances in downgradient groundwater of PBTVs and 2L or IMAC are considered a primary focus. Additionally, NCDEQ requested that hexavalent chromium be included as a COI at each CAMA-related site due to public interest and receptor wells. Molybdenum and strontium do not have 2L or IMACs established; however, these constituents are considered potential COIs with regards to CCR and are evaluated as potential COIs for the Site at the request of NCDEQ. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 10-12 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx The following constituents do not have a 2L standard, IMAC, or Federal MCL established: alkalinity, aluminum, bicarbonate, calcium, carbonate, magnesium, methane, potassium, sodium, sulfide, and TOC. Results from these constituents are useful in comparing water conditions throughout the Site. For example calcium is listed as a constituent for detection monitoring in Appendix III to 40 CFR Part 257. Although these constituents will be used to compare and understand groundwater quality conditions at the Site, because there are no associated 2L, IMACs, or MCLs, these constituents are not evaluated as potential COIs for the Site. Additional Requirements 10.3.3 NCDEQ requested that figures be included in the CSA that depict groundwater analytical results for the constituents in 40 CFR 257, Appendix III detection monitoring and 40 CFR 257, Appendix IV assessment monitoring (CCR Rule) (USEPA, 2015). Detection monitoring constituents in 40 CFR 257 Appendix III are: Boron Calcium Chloride Fluoride (limited historical data at this Site, not on assessment constituent list) pH Sulfate TDS Constituents for assessment monitoring listed in 40 CFR 257 Appendix IV include: Antimony Arsenic Barium Beryllium Cadmium Chromium Cobalt 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 10-13 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx Fluoride (limited historical data at this Site, not on assessment constituent list) Lead Lithium (not analyzed) Mercury Molybdenum Selenium Thallium Radium 226 and 228 combined Aluminum, copper, iron, manganese, and sulfide were originally included in the Appendix IV constituents in the draft rule; USEPA removed these constituents in the final rule. Therefore, these constituents are not included in the listing above; however, they are included as part of the current Interim Monitoring Plan (IMP; Section 15.3). NCDEQ requested that vanadium be included as a COI. Mayo Plant COIs 10.3.4 Exceedances of comparative values, the distribution of constituents in relation to the ash basin, comparison with background concentrations, co-occurrence with CCR indicator constituents such as boron, and likely migration directions based on groundwater flow direction are considered in determination of groundwater COIs. A constituent exceedance in an outlying area with no co-occurrence of boron or similar CCR-related constituent would likely not be considered a reason to list the constituent as a COI. A constituent exceedance based on a single sampling event when previous results indicate a concentration trend below comparative values would likely not indicate inclusion as a COI. Based on site- specific conditions, observations, and findings, the following list of COIs has been developed for the Mayo Plant: Arsenic Barium Boron Chromium (total) Chromium (hexavalent) Cobalt Iron 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 10-14 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx Manganese Molybdenum pH Strontium Sulfate TDS Vanadium Table 10-2 lists the COIs and other constituents at the Mayo Site along with the established PBTVs and associated 2L/IMACs. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 11-1 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx 11.0 HYDROGEOLOGICAL INVESTIGATION Results from the hydrogeological assessment of the Mayo Site, summarized in this section, are primary components of the SCM.3 Plume physical and chemical characterization is detailed below for each groundwater COI. The horizontal and vertical extent of constituent concentrations is presented on isoconcentration maps and cross sections. These descriptions and depictions are based primarily on the most recent comprehensive groundwater sampling event (March-April 2017). 11.1 Plume Physical and Chemical Characterization Plume Physical Characterization 11.1.1 The groundwater plume is defined as locations (in three-dimensional space) where groundwater quality is impacted by the ash basin. Other COIs (defined in Section 10.0) are used to help refine the extent and degree to which areas are impacted by groundwater from the ash basin. The comprehensive groundwater data table (Appendix B) and an understanding of groundwater flow dynamics and direction (Section 6.2.3, Figures 6-4 to 6-11) were used to define the horizontal and vertical extent of the plume. As discussed in Section 13.2 (Geochemical Modeling), not all constituents with PBTV exceedances can be attributed to the ash basin. Naturally occurring groundwater contains varying concentrations of alkalinity, aluminum, bicarbonate, cadmium, carbonate, copper, lead, magnesium, methane, nickel, potassium, sodium, TOC, and zinc. Sporadic and low-concentration exceedances of these constituents in the groundwater data do not necessarily demonstrate horizontal or vertical distribution in groundwater that indicates impact from the ash basin. Isoconcentration Maps The horizontal extent of the plume in each flow unit is interpreted in concentration isopleth maps (Figures 11-1 to 11-45). These maps use valid 3 Pursuant to the CCR rule, owners and operators of CCR units must install the required groundwater monitoring system; develop the required groundwater sampling and analysis program to include selection of the statistical procedures to be used for evaluating groundwater monitoring data; and begin detection monitoring, which requires owners and operators to have a minimum of eight samples for each well and begin evaluating groundwater monitoring data for statistically significant increases over background levels for the constituents listed in Appendix III of 40 C.F.R. Part 257. These data need not be posted to Duke Energy’s publicly accessible Internet site until such time the annual groundwater monitoring and corrective action report required under the CCR rule becomes due. Although a portion of these data was utilized in this assessment for refinement of constituent distribution, these data are not included in this report because it was not public information as of the date of its completion. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 11-2 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx groundwater analytical data to spatially and visually define areas where groundwater concentrations are above the respective constituent PBTV and/or 2L/IMAC. The leading edge of the plume, the furthest downgradient edge, is represented by groundwater concentrations in the wells in each flow unit. In the bedrock flow unit, boron is detected in downgradient well CW-2D (located near the northern property line). In bedrock well MW-16BR, located approximately 1,000 feet downgradient of CW-2D and the property line, boron is not detected. The leading edge of the bedrock plume is interpreted to be near the northern property line. Similarly, in the transition zone unit, groundwater from monitoring well CW-2 contains detectable concentrations of boron while MW- 16D, located about 1,000 feet downgradient, does not. Surficial well MW-3 is located several hundred feet south of the property line between the property line and the ash basin dam. MW-16S is located downgradient approximately 1,500 feet from MW-3. Both wells are screened in alluvium. Boron has been detected in both wells. Figures 11-7 to 11-9 depict the horizontal extent of boron in downgradient groundwater. The background contour line in the surficial unit encompasses the perimeter of the ash basin, extends beyond the Duke Energy property boundary to the north, and ends just south of the state line. In the transition zone, the background contour line mimics the general shape of the plume in the surficial unit except in the downgradient direction where the extent of impact is between the property line and the MW-16 well cluster. In bedrock, the background contour encompasses only the northeast half of the ash basin and also extends just northeast of the property line, but the background contour remains south of the MW-16 well cluster. As described in Section 6.0, there is no hydrogeologic confining unit at Mayo; therefore, under these unconfined conditions, groundwater moves freely across each unit. Concentration versus Distance Plots Figure 11-46 to 11-48 depicts concentration versus distance graphs from the source along the plume centerline for COIs. While PBTV values could not be distinguished on these graphs because values differ by flow unit, the graphs show constituent concentrations in source areas and downgradient and aid in understanding plume distribution. Concentrations of each COI represent March- April 2017 conditions. The wells used are consistent for each constituent 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 11-3 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx represented. The graphs demonstrate that COI concentrations decrease from the source area to downgradient locations. Vertical Extent Cross-Sections The vertical extent of the plume down the centerline is depicted in the cross- sectional views of the Site (Figures 11-49 to 11-63). Cross-section C-C’ is a more detailed and focused depiction of cross-section B-B’, and focuses on the ash basin and directly downgradient. Wells along the centerline plume flow path are depicted. These wells represent source area and downgradient locations relative to the ash basin. COIs have been contoured in the cross-sectional depictions. Constituent isopleths reflect values above the PBTV and the 2L/IMAC standard, as applicable. The surficial and transition zone flow units at Mayo are not vertically extensive (less than 10 feet thick for surficial; generally less than 30 feet thick for transition zone). The well screens in the CAMA wells accurately monitor groundwater conditions and impact to the groundwater flow zones for the surficial and transition zone units. ABMW-4D is in the transition zone below the ash basin and contains detectable concentrations of boron (between 3,470 – 5,090 ug/L). Other COI constituent concentrations also exceed 2L/IMAC and PBTV values; therefore, the transition zone beneath the basin appears to be impacted. The vertical extent of the plume is best represented by groundwater concentrations in bedrock wells beneath and downgradient of the ash basin. Deep bedrock well ABMW-2BRL contains no boron or manganese concentrations above 2L or PBTV, respectively. ABMW-2BR and ABMW-4BR are shallower bedrock wells beneath the ash basin, and groundwater from these wells has the same general absence of CCR-related constituents at concentrations that exceed PBTVs or the 2L/IMAC. Upward vertical gradients as groundwater from the west, south, and east recharge the groundwater beneath the basin reduce the potential for downward migration of COIs into bedrock. As groundwater under the ash basin flows northeast toward the ash basin dam, the hydraulic impact of the ash basin dam and the hydraulic head exerted by the ash basin water forces groundwater downward into the bedrock, which increases hydraulic pressure in the bedrock aquifer. In general, the pressures in the bedrock just downgradient of the base of the dam become greater than in the 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 11-4 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx transition zone or surficial aquifers resulting in upward vertical gradients. A strong upward gradient exists along the western end of the dam and gradually changes (eastward) to a downward or neutral gradient in the former Crutchfield Branch valley. This is expected as discharging groundwater and surface water reach hydraulic equilibrium. As groundwater migrates in the downgradient direction, groundwater enters the system from upgradient recharge areas to the west and east. In summary, the horizontal and vertical extent of the plume has been defined. Monitoring wells across the Site are appropriately placed and screened to the correct elevations to monitor groundwater quality. Monitoring wells installed for other regulatory programs 4 have added additional details about the orientation and extent of the downgradient plume and have helped refine an understanding of the vertical and horizontal distribution of the plume. Plume Chemical Characterization 11.1.2 Plume chemical characterization is detailed below for each COI. Analytical results are based on the March-April 2017 groundwater sampling event. The range of detected concentrations is presented with the number of detections for the sampling event. Descriptions of the COIs identified for the Mayo Site are also provided. PBTVs and 2L/IMACs are included in Appendix B, Table 1. Pore water (source) concentrations are discussed in Section 3.0. Arsenic Detected Range: 1.06 µg/L – 26.9 µg/L; Number of Detections/Total Samples: 7/33 Concentrations in 5 samples (1 surficial; 4 bedrock) exceeded the PBTV. Concentrations in 1 sample (transition zone beneath ash basin) exceeded the 2L of 1 µg/L. 4 Pursuant to the CCR rule, owners and operators of CCR units must install the required groundwater monitoring system; develop the required groundwater sampling and analysis program to include selection of the statistical procedures to be used for evaluating groundwater monitoring data; and begin detection monitoring, which requires owners and operators to have a minimum of eight samples for each well and begin evaluating groundwater monitoring data for statistically significant increases over background levels for the constituents listed in Appendix III of 40 C.F.R. Part 257. These data need not be posted to Duke Energy’s publicly accessible Internet site until such time the annual groundwater monitoring and corrective action report required under the CCR rule becomes due. Although a portion of these data was utilized in this assessment for refinement of constituent distribution, these data are not included in this report because it was not public information as of the date of its completion 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 11-5 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx Arsenic is a trace element in the crust, with estimated concentrations ranging from less than one mg/kg in mafic igneous rocks to 13 mg/kg in clay-rich rocks (Parker, 1967). It occurs in multiple valence states (As5+, As3+, and As3-). Arsenic in coal occurs primarily in pyrite (iron sulfide, with arsenic replacing iron in the crystal structure) (Finkelman, 1995). Arsenic condenses on fly ash as arsenate (As5+) (Goodarzi, Huggins, & Sanei, 2008). Leaching tests on ash indicate that trace quantities of up to 50 percent of the arsenic present can be leached. In addition to the solubility of the source, the concentration of calcium and presence of oxides appear to limit the mobility of arsenic (Izquierdo & Querol, 2012). The USEPA estimates that the amount of natural arsenic released into the air as dust from the soil is approximately equal to the amount of arsenic released by all human activities (EPRI, 2008a). Barium Detected Range: 6 µg/L – 890 µg/L; Number of Detections/Total Samples: 38/33 Concentrations in 7 samples (3 surficial; 2 transition zone; 2 bedrock) exceeded the PBTV. Concentrations in 1 sample (transition zone beneath the ash basin) exceeded the 2L of 700 µg/L. Historic barium detected concentrations align with the primary path of groundwater flow transect including historic exceedances greater than the PBTV and 2L, including in downgradient surficial and transition zone groundwater. Barium is a naturally occurring component of minerals that are found in small, but widely distributed amounts in the earth’s crust (Kunesh, 1978); (Miner, 1969). Two forms of barium, barium sulfate (barite) and barium carbonate (witherite) are often found in nature as ore deposits. Barium enters the environment naturally through the weathering of rocks and minerals. Anthropogenic releases are associated primarily with industrial processes. Barium is sometimes found naturally in drinking water and food. However, because the dominant naturally occurring barium compounds (barium sulfate and barium carbonate) have a low to moderate solubility in water under most conditions, the amount of barium found in drinking water is typically small. Barium compounds such as barium acetate, barium chloride, barium hydroxide, 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 11-6 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx barium nitrate, and barium sulfide dissolve more easily in water than barium sulfate and barium carbonate, but because they are not commonly found in nature, the latter two compounds do not usually occur in drinking water unless the water is contaminated by barium compounds that are released from waste sites (EPRI, 2008b). Barium is naturally released into the air by soils as they erode and is released into the soil and water by eroding rocks. Barium released into the air by human activities comes mainly from barium mines, metal production facilities, and industrial boilers that burn coal and oil (EPRI, 2008b). The leachability of barium has been found to be relatively independent of pH but is controlled instead by the presence of calcium, with which it competes for sulfate (Fruchter, Rai, & Zachara, 1990). In an overview of leachability studies found in the International Journal of Coal Geology, the mobility of barium typically ranged from 0.02 percent to 2 percent (Izquierdo & Querol, 2012). Regional metamorphic grade greenschist to upper amphibolite in the Piedmont’s King’s Mountain Belt contains deposits of barium sulfate (barite). Barium is especially common as concretions and vein fillings in limestone and dolostone, which are not common geologic facies in North Carolina; however, at various times in the past century and a half, the Carolinas have been major producers of barite (USEPA, 2017a) . In a statistical summary of groundwater quality in North Carolina, the Superfund Research Program at the University of North Carolina (UNC) analyzed 1,898 private well water samples in Gaston and Mecklenburg Counties. The samples were tested by the North Carolina State Laboratory of Public Health from 1998-2012. This study found an average barium concentration of 50 µg/L. No samples exceeded the 2,000 µg/L Primary Maximum Contaminant Level (PMCL) for barium (NCDHHS, 2010a). Boron Detected Range: 179 µg/L – 3,470 µg/L; Number of Detections/Total Samples: 6/33 Concentrations in 6 samples (3 surficial; 2 transition zone; 1 bedrock) exceeded the PBTV. Concentrations in 4 samples exceeded the 2L of 700 µg/L. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 11-7 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx Boron exceeds the PBTV and 2L in surficial and transition zone groundwater beneath the ash basin. Boron is a trace element in the crust, with estimated concentrations ranging from as little as 1 mg/kg in mafic igneous rocks to hundreds of milligrams per kilogram in clay rich rocks (Parker, 1967). It occurs only in the trivalent form (B3+) and is concentrated in sedimentary rocks (Urey & Mem, 1953). This observation indicates that a mechanism exists to concentrate boron in minerals because the oceans could dissolve all of the boron estimated to be present in the crust (Fleet, 1965). Fleet (1965) presents both biogenic and mineralogical processes to account for the preferential concentration of boron in the crust. Boron is a micronutrient (Goldberg, 1997) that is concentrated in plant tissue, including the plants from which coal formed. While boron is relatively abundant on the earth’s surface, boron and boron compounds are relatively rare in all geological provinces of North Carolina. Natural sources of boron in the environment include volatilization from seawater, geothermal vents, and weathering of clay-rich sedimentary rocks. Total contributions from anthropogenic sources are less than contributions from natural sources. Anthropogenic sources of boron include agriculture, refuse, coal and oil burning power plants, by-products of glass manufacturing, and sewage and sludge disposal (EPRI, 2005). Because boron is associated with the carbon (fuel) in coal, it tends to volatilize during combustion and subsequently condense onto fly ash as a soluble borate salt (Dudas, 1981). Boron leaches readily (up to 50 percent of total present) and rapidly from fly ash (Cox, Lundquist, Przyjazny, & Schmulbach, 1978). Boron is considered a marker COI for coal ash because boron is rarely associated with other types of industrial waste products. Boron is the primary component of a few minerals including tourmaline, a rare gem mineral that forms under high temperature and pressure (Hurlbut, 1971). The remaining common boron minerals, including borax that was mined in the Mojave desert, in Boron, California, form from the evaporation of seawater in deposits known as evaporites. For this reason, boron mobilized into the environment will remain in solution until incorporation into plant tissue or adsorption by a mineral. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 11-8 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx Fleet (1965) describes sorption of boron by clays as a two-step process. Boron in solution is likely to be in the form of the borate ion (B(OH)4-). The initial sorption occurs onto a charged surface. Observations that boron does not tend to desorb from clays indicates that it migrates rapidly into the crystal structure, most likely in substitution for aluminum. Goldberg et al. (1996) determined that boron sorption sites on clays appear to be specific to boron. For this reason, there is no need to correct for competition for sorption sites by other anions in transport models (Goldberg, Forster, Lesch, & Heick, 1996). Goldberg (1997) lists aluminum and iron oxides, magnesium hydroxide, clay minerals, calcium carbonate (limestone), and organic matter as important sorption surfaces in soils (Goldberg, 1997). Boron sorption on oxides is diminished by competition from numerous anions. Boron solubility in groundwater is controlled by adsorption reactions rather than by mineral solubility. Goldberg concludes that chemical models can effectively replicate boron adsorption data over changing conditions of boron concentration, pH, and ionic strength. Chromium Detected Range: 1.3 µg/L – 9.58 µg/L; Number of Detections/Total Samples: 8/33 Concentrations in 1 sample (transition zone) exceeded the PBTV. No exceedances of the 2L of 10 µg/L were observed. Chromium is a blue-white metal found naturally occurring in combination with other substances. It occurs in rocks, soils, plants, and volcanic dust and gases (EPRI, 2008c). Background concentrations of chromium in groundwater generally vary according to the media in which they occur. Most chromium concentrations in groundwater are low averaging less than 1.0 µg/L worldwide. Chromium tends to occur in higher concentrations in felsic igneous rocks (such as granite and metagranite) and ultramafic igneous rocks; however, it is not a major component of the igneous or metamorphic rocks found in the North Carolina Piedmont or the Blue Ridge (Chapman, Cravotta, III, Szabo, & Lindsey, 2013) In a statistical summary of groundwater quality in North Carolina, the Superfund Research Program at UNC analyzed 1,898 private well water samples in Gaston and Mecklenburg Counties. The samples were tested by the North 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 11-9 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx Carolina State Laboratory of Public Health from 1998 to 2012. The average chromium concentrations were 5.1µg/L and 5.2 µg/L in Gaston and Mecklenburg Counties respectively. Hexavalent Chromium Detected Range: 0.028 µg/L – 4.4 µg/L; Number of Detections/Total Samples: 18/33 Concentrations in 4 samples (2 surficial; 2 bedrock) exceeded the PBTV. No exceedances of the 2L of 10 µg/L were observed. Chromium can also occur in the +III oxidation state, depending on pH and redox conditions. Cr (VI) is the dominant form of chromium in shallow aquifers where aerobic conditions exist. Cr(VI) can be reduced to Cr(III) by soil organic matter, S2- and Fe2+ ions under anaerobic conditions often encountered in deeper groundwater. Major Cr(VI) species include chromate (CrO4 2-) and dichromate (Cr2O7 2-) which precipitate readily in the presence of metal cations (especially Ba2+, Pb2+, and Ag+). Chromate and dichromate also adsorb on soil surfaces, especially iron and aluminum oxides. Cr(III) is the dominant form of chromium at low pH. Chromium mobility depends on sorption characteristics of the soil, including clay content, iron oxide content, and the amount of organic matter present. Chromium can be transported by surface runoff to surface waters in its soluble or precipitated form. Soluble and unadsorbed chromium complexes can leach from soil into groundwater. The leachability of Cr(VI) increases as soil pH increases. Most of chromium released into natural waters is particle associated, however, and is ultimately deposited into the sediment (Smith, Means, Chen, & others, 1995). Cobalt Detected Range: 1.59 µg/L – 5.9 µg/L; Number of Detections/Total Samples: 4/33 Concentrations in 4 samples (2 surficial; 1 transition zone; 1 bedrock) exceeded the PBTV and IMAC of 1 µg/L. Cobalt is a base metal that exhibits geochemical properties similar to iron and manganese, occurring as a divalent and trivalent ion. Cobalt can also occur as Co1-. In terms of distribution in the crust, all three metals exhibit a strong affinity 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 11-10 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx for mafic igneous and volcanic rocks and deep-sea clays (Parker, 1967). Cobalt occurs in clay minerals and substitutes into the pyrite crystal structure. There is also evidence that it is organically bound in coal (Finkelman, 1995). Izquierdo and Querol (2012) report limited leaching of cobalt from coal, attributing this observation to incorporation into iron oxide minerals. The concentration of cobalt in surface and groundwater in the United States is generally low— between 1 and 10 parts of cobalt in 1 billion parts of water (parts per billion; ppb) in populated areas. The concentration may be hundreds or thousands times higher in areas that are rich in cobalt containing minerals or in areas near mining or smelting operations. In most drinking water, cobalt levels are less than 1 to 2 ppb (USGS, 1973). Cobalt is compared to IMAC since no 2L standard has been established for this constituent by NCDEQ. Iron Detected Range: 11 µg/L – 58,800 µg/L; Number of Detections/Total Samples: 28/33 Concentrations in 5 samples (2 surficial; 2 transition zone; 1 bedrock) exceeded the PBTV. Concentrations in 17 samples exceeded the 2L of 300 µg/L. Iron exceeds the PBTV and 2L in surficial and transition zone groundwater beneath the ash basin. Iron is a naturally occurring element that may be present in groundwater from the erosion of natural deposits (NCDHHS, 2010b). A 2015 study by NCDEQ (Summary of North Carolina Surface Water Quality Standards 2007-2014) found that while concentrations vary regionally, “iron occurs naturally at significant concentrations in the groundwaters of NC,” with a statewide average concentration of 1,320 µg/L. Iron is estimated to be the fourth most abundant element in the Earth’s crust at approximately 5 percent by weight (Parker, 1967). Only Oxygen (46.60 weight percent), silicon (27.72 weight percent), and aluminum (8.13 weight percent) occur in higher concentrations. Iron occurs in divalent (ferrous, Fe2+), trivalent (ferric, Fe3+), hexavalent (Fe6+), and Fe2- oxidation states. Iron is a common mineral-forming element, occurring primarily in mafic (dark colored) minerals including micas, pyrite (iron disulfide), and hematite (iron oxide), as well as in reddish-colored clay minerals. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 11-11 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx Clay minerals and pyrite are common impurities in coal. Under combustion conditions in a coal-fired boiler, clay minerals would be dehydrated to mullite or gibbsite, possibly liberating iron, and pyrite would oxidize to hematite or magnesioferrite. Research summarized by Izquierdo and Querol (2012) indicates that iron leaching from coal ash is on the order of one percent of the total iron present due to the low pH required to solubilize iron minerals. Despite the low apparent mobilization percentage, iron is often one of the COIs detected in the highest concentrations in ash pore water. Ferric iron is soluble at pH less than 2 at typical surface conditions (25°C and one atmosphere total pressure (Schmidt, 1962). For this reason, dissolved iron in surficial waters is typically oxidized to the trivalent state resulting in formation of ferric iron oxide flocculation that exhibits a characteristic reddish tint. Manganese Detected Range: 6 µg/L – 6,960 µg/L; Number of Detections/Total Samples: 29/33 Concentrations in 12 samples (3 surficial; 2 transition zone; 7 bedrock) exceeded the PBTV. Concentrations in 25 samples exceeded the 2L of 50 µg/L. Manganese exceeds the PBTV and 2L in surficial and transition zone groundwater beneath the ash basin. Manganese is a naturally occurring silvery-gray transition metal that resembles iron but is more brittle and is not magnetic. It is found in combination with iron, oxygen, sulfur, or chlorine to form manganese compounds. High manganese concentrations are associated with silty soils, and sedimentary, unconsolidated, or weathered lithologic unit and low concentrations are associated with non- weathered igneous bedrock and soils with high hydraulic conductivity (Gillespie, 2013), (Polizzotto, et al., 2015). Manganese is most readily released to the groundwater through the weathering of mafic or siliceous rocks (Gillespie, 2013). When manganese-bearing minerals in saprolite, such as biotite, are exposed to acidic weathering, the metal can be liberated from the host mineral and released to groundwater. It then migrates through pre-existing fractures during the movement of groundwater through bedrock. If this aqueous-phase manganese is exposed to higher pH in the groundwater system, it will precipitate out of solution. This results in preferential pathways becoming “coated” in manganese oxides and introduces a concentrated source of 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 11-12 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx manganese into groundwater (Gillespie, 2013). Manganese(II) in suspension of silt or clay is commonly oxidized by microorganisms present in soil, leading to the precipitation of manganese minerals (ATSDR, 2012). Roughly 40 percent to 50 percent of North Carolina wells have manganese concentrations higher than the state drinking water standard (Gillespie, 2013). Concentrations are spatially variable throughout the state, ranging from less than 0.01 mg/L to more than 2 mg/L. This range of values reflects naturally derived concentrations of the constituent and is largely dependent on the bedrock’s mineralogy and extent of weathering (Gillespie, 2013). Manganese is estimated to be the 12th most abundant element in the crust (0.100 weight percentage, (Parker, 1967)). Manganese exhibits geochemical properties similar to iron with Mn7+, Mn6+, Mn4+, Mn3+, Mn2+, and Mn1- oxidation states. Manganese substitutes for iron in many minerals. Similar to iron, manganese leaching from coal ash is limited to less than 10 percent of the total manganese present due to the low pH required to solubilize manganese minerals (Izquierdo & Querol, 2012). Despite the low apparent mobilization percentage, manganese can be detected in relatively high concentrations in ash pore water. Molybdenum Detected Range: 1.02 µg/L – 1,560 µg/L; Number of Detections/Total Samples: 22/37 Concentrations in 8 samples exceeded the PBTV. Molybdenum exceeds the PBTV in pore water and transition zone groundwater beneath the ash basin. Molybdenum detected in groundwater downgradient of the ash basin including exceedances of PBTV in transition zone and bedrock groundwater, including off-site. Molybdenum is a trace element that exists predominantly as Mo(IV) and Mo(VI). As a free metal, it is silvery gray, although it does not occur in this form in nature. It is mined for use in alloys. Molybdenum commonly forms oxyanions in groundwater that are affected by redox and pH (Ayotte, Gronbert, & Apodaca, 2011). Molybdenum has been observed to leach less from coal cleaning rejects in acidic than neutral conditions, unlike many other metals (Jones & Ruppert, 2017). Molybdenum has been shown to become more mobile in procedures that use 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 11-13 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx deionized water as a leachant, which may be similar to actual disposal conditions unlike many other coal ash elements that are more mobile when subjected to weak acid (Jones & Ruppert, 2017). Strontium Detected Range: 27 µg/L – 2,320 µg/L; Number of Detections/Total Samples: 33/33 Concentrations in 12 samples (4 surficial; 2 transition zone; 6 bedrock) exceeded the PBTV. Strontium does not have a 2L or IMAC. Strontium exceeds the PBTV in surficial, transition zone, and bedrock groundwater beneath the ash basin. Strontium is a soft silver-yellow alkaline earth metal. It is highly chemically reactive and forms a dark oxide layer when it interacts with air. It is chemically similar to Ca and replaces Ca or K in igneous rocks in minor amounts. Strontium is generally present in low concentrations in surface waters but may exist in higher concentrations in some groundwater (Hem, 1985). Strontium is present as a minor coal and coal ash constituent. Strontium has been observed to leach from coal cleaning rejects more in neutral conditions than acidic, unlike many other metals (Jones & Ruppert, 2017). It has been shown to behave conservatively in surface waters downstream of coal plants (Ruhl, et al., 2012). Sulfate Detected Range: 0.27 µg/L – 62 µg/L; Number of Detections/Total Samples: 37/37 Concentrations in 17 samples (2 surficial; 7 transition zone; 8 bedrock) exceeded the PBTV. No exceedances of the 2L of 250 µg/L were observed. Sulfate is a naturally occurring substance found in minerals, soil, and rocks. It is present in ambient air, groundwater, plants, and food. Primary natural sources of sulfate include atmospheric deposition, sulfate mineral dissolution, and sulfide mineral oxidation. The principal commercial use of sulfate is in the chemical industry. Sulfate is discharged into water in industrial wastes and 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 11-14 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx through atmospheric deposition (USEPA, 2003). Anthropogenic sources include coal mines, power plants, phosphate refineries, and metallurgical refineries. Sulfate has a Secondary Maximum Contaminant Level (SMCL), and no enforceable maximum concentration set by the USEPA. Ingestion of water with high concentrations of sulfate may be associated with diarrhea, particularly in susceptible populations, such as infants and transients (USEPA, 2012). However, adults generally become accustomed to high sulfate concentrations after a few days. It is estimated that about 3 percent of the public drinking water systems in the United States may have sulfate concentrations of 250 mg/L or greater (Miao, Brusseau, Carroll, & others, 2012). Sulfate is on the list of enforced regulated contaminates that may cause cosmetic effects or aesthetic effects in drinking water (USEPA, 2017a). TDS Detected Range: 51 mg/L – 730 mg/L; Number of Detections/Total Samples: 33/33 Concentrations in 12 samples (3 surficial; 2 transition zone; 7 bedrock) exceeded the PBTV. Concentrations in 6 samples exceeded the 2L of 500 mg/L. TDS exceeds the PBTV in surficial, transition zone, and bedrock groundwater beneath the ash basin. Groundwater contains a wide variety of dissolved inorganic constituents as a result of chemical and biochemical interactions between the groundwater and the elements in the soil and rock through which it passes. TDS mainly consist of cation and anion particles (e.g., calcium, chlorides, nitrate, phosphorus, iron, sulfur, and others) that can pass through a 2 micron filter (USEPA, 1997). TDS is therefore a measure of the total amount of dissolved ions in the water, but does not identify specific constituents or explain the nature of ion relationships. TDS concentrations in groundwater can vary over many orders of magnitude and generally range from 0 – 1,000,000 µg/L. The ions listed below are referred to as the major ions as they make up more than 90 percent of the TDS in groundwater. TDS concentrations resulting from these constituents are commonly greater than 5,000 µg/L (Freeze & Cherry, 1979). 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 11-15 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx Sodium (Na+) Magnesium (Mg2+) Calcium (Ca2+) Chloride (Cl-) Bicarbonate (HCO3-) Sulfate (SO4 2-) Minor ions in groundwater include: boron, nitrate, carbonate, potassium, fluoride, strontium, and iron. TDS concentrations resulting from minor ions typically range between 10 – 1,000 µg/L (Freeze & Cherry, 1979). Trace constituents make up an even smaller portion of TDS in groundwater and include: aluminum, antimony, arsenic, barium, beryllium, cadmium, chromium, cobalt, lead, manganese, nickel, selenium, thallium, vanadium, and zinc among others. TDS concentrations resulting from trace constituents are typically less than 100 µg/L (Freeze & Cherry, 1979). In some cases, contributions from anthropogenic sources can cause some of the elements listed as minor or trace constituents to occur as contaminants at concentration levels that are orders of magnitude above the normal ranges indicated above. TDS in water supplies originate from natural sources, sewage, urban and agricultural run-off, and industrial wastewater. Salts used for road de-icing can also contribute to the TDS loading of water supplies. Concentrations of TDS from natural sources have been found to vary from less than 30 mg/L to as much as 6,000 mg/L. Water containing more than 2,000 – 3,000 mg/L TDS is generally too salty to drink (the TDS of seawater is approximately 35,000 mg/L) (Freeze & Cherry, 1979). Reliable data on possible health effects associated with the ingestion of TDS in drinking water are not available (WHO, 1996). TDS is on the list of “National Secondary Drinking Water Regulations” (NSDWRs) which are non-enforced regulated contaminates that may cause cosmetic effects or aesthetic effects in drinking water (USEPA, 2017b). Vanadium Detected Range: 0.308 µg/L – 56.7 µg/L; Number of Detections/Total Samples: 22/37 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 11-16 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx Concentrations in 2 samples (1 surficial; 1 transition zone) exceeded the PBTV. Concentrations in 17 samples exceeded the IMAC of 0.3 µg/L. Vanadium exceeds the PBTV and IMAC in transition zone, and bedrock groundwater beneath the ash basin. Vanadium is estimated to be the 22nd most abundant element in the crust (0.011 weight percent, (Parker, 1967). Vanadium occurs in four oxidation states (V5+, V4+, V3+, and V2+). It is a common trace element in both clay minerals and plant material. The National Uranium Resource Evaluation (NURE) program was initiated by the Atomic Energy Commission in 1973 with a primary goal of identifying uranium resources in the United States (Smith S. M., 2016). The Hydrogeochemical and Stream Sediment Reconnaissance program (initiated in 1975) was one component of NURE. Planned systematic sampling of the entire United States began in 1976 under the responsibility of four Department of Energy national laboratories. Samples were collected from 5,178 wells across North Carolina. Of these, the concentration of vanadium was equal to or higher that the former IMAC of 0.0003 mg/L in 1,388 well samples (27 percent). pH Detected Range: 4.9 – 7.9 Concentrations in 8 samples (1 surficial; 3 transition zone; 3 bedrock) exceeded the PBTV. Concentrations in 12 samples exceeded the 2L. The pH scale is used to measure acidity or alkalinity. A pH value of 7 indicates neutral water. A value lower than the USEPA-established SMCL range (<6.5 Standard Units) is associated with a bitter, metallic tasting water, and corrosion. A value higher than the SMCL range (>8.5 Standard Units) is associated with a slippery feel, soda taste, and deposits in the water (USEPA, 2017b). In a statistical summary of groundwater quality in North Carolina, the Superfund Research Program at UNC analyzed 618 private well water samples for pH in Cleveland and Rutherford Counties. The samples were analyzed by the 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 11-17 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx North Carolina State Laboratory of Public Health from 1998 – 2012. This study found that 16.9 percent of wells in Cleveland County and 20.3 percent of wells in Rutherford County had a pH result outside of the USEPA’s SMCL range. 11.2 Pending Investigation(s) Additional metal oxy-hydroxide phases of iron (HFO) and aluminum (HAO) data are needed to support geochemical modeling conducted as part of the CAP. Soil and rock samples from previously installed borings or from additionally drilled boreholes along the primary groundwater flow transect will be used. The samples will be located: Background/Upgradient Directly beneath ash basin Downgradient location, north of the ash basin The samples will be collected at vertical intervals that coincide with nearby well screen elevations. Analysis results of collected samples will be used to improve input parameters for the updated geochemical model. In accordance with 15A NCAC 02L.0106( k)(5) and (l)(6), the CAP may include an evaluation of whether groundwater migrating downgradient of the ash basin may have contaminant concentrations that would result in violations of standards for surface water. One means of accomplishing this objective is to collect surface water samples to document existing conditions. Another method will be to conduct groundwater to surface water modeling. It is anticipated that documenting current conditions through the collection of the additional surface water samples in Crutchfield Branch will be coordinated with NCDEQ guidance. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 12-1 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx 12.0 RISK ASSESSMENT A baseline human health and ecological risk assessment was performed as a component of CAP, Part 2 (SynTerra, 2016a). The 2016 risk assessment characterized potential risks to humans and wildlife exposed to coal ash constituents present in environmental media for the purpose in aiding corrective action decisions. Implementation of corrective action is intended to achieve future Site conditions protective of human health and the environment, as required by CAMA. This update to the risk assessment evaluates groundwater and surface water results collected since the 2016 risk assessment (November 2015 to June 2017) in order to confirm or update risk conclusions in support of remedial action. Data used in the 2016 risk assessment included groundwater samples from March 11, 2015 through September 11, 2015, surface water samples from August 27, 2014 through October 13, 2015, and AOW water samples from May 12, 2015 through October 12, 2015. AOW soil and sediment samples were collected May 12 and 13, 2015. This risk assessment update uses sampling locations described in Attachment A of the 2016 risk assessment (SynTerra, 2016a). As previously noted, AOW locations are outside the scope of this risk assessment because AOWs, wastewater, and wastewater conveyances (effluent channels) are evaluated and governed wholly separate in accordance with the NPDES Program administered by NCDEQ DWR. This process is on-going in a parallel effort to the CSA and subject to change. No new sediment or soil samples, other than for background evaluations, have been collected that are applicable to the risk assessment; therefore, risk estimates associated with those media have not been re-evaluated. As part of the 2016 risk assessment, human health and ecological conceptual site models (CSMs) were developed to guide identification of exposure pathways, exposure routes, and potential receptors for evaluation in the risk assessment. The CSMs (CAP, Part 2; Figures 2-2 and 2-4) describe the sources and potential migration pathways through which groundwater beneath the ash basin may have transported coal ash-derived constituents to other environmental media (receiving media) and, in turn, to potential human and ecological receptors. Exposure scenarios and exposure areas were presented in detail in the 2016 CAP, Part 2 risk assessment. This risk assessment update included the following: Identification of maximum constituent concentrations for groundwater and surface water; 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 12-2 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx Inclusion of new groundwater and surface water data to derive overall average constituent concentrations for exposure areas; Comparison of new maximum constituent concentrations to the risk assessment human health and ecological screening values; Comparison of new maximum constituent concentrations to human health Risk- Based Concentrations (RBC); and Incorporation of new maximum constituent concentrations into wildlife Average Daily Dose (ADD) calculations for comparison to ecological Toxicity Reference Values (TRVs). Evaluation of new groundwater and surface water data and the influence on the 2016 risk assessment conclusions are summarized below by exposure areas (Figures 12-1 and 12-2) at the Mayo Plant. 12.1 Human Health Screening Summary On-site Groundwater – Surficial Aquifer Groundwater sample locations used in the human health assessment of the surficial aquifer include: ABMW-3S, MW-3 and MW-16S. These wells were evaluated because they represent the potential trespasser/worker exposure area as determined in the 2016 risk assessment. Groundwater analytical results are included in Appendix B, Table 1. Dissolved thallium (0.45 micron filter) detected at a concentration of 0.227 µg/L in sample ABMW-3S collected on December 1, 2015 exceeded the human health screening value of 0.2µg/L. This reported detection is likely anomalous, as it was the only thallium detection in the surficial aquifer dataset of 23 samples, and is less than the total thallium concentration (<0.2 µg/L) analyzed in the same sample. No potential risks to humans exposed to groundwater from the surficial aquifer were identified. On-site Groundwater – Bedrock and Transition Zone Aquifer Groundwater sample locations used in the human health assessment of the bedrock and transition zone aquifer include: ABMW-2BR, ABMW-4BR, ABMW-4D, CW-1/1D through CW-6, MW-2, MW-3BR, MW-4, MW-7D and MW-16D. Data for the groundwater sample locations are included in Appendix B, Table 1. One thallium (total) detection of 0.21 µg/L collected on December 1, 2015 in sample ABMW-2BR exceeded the human health screening value of 0.2 µg/L. This reported detection is likely anomalous, as it was the only thallium detection in the bedrock and transitional 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 12-3 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx zone aquifer dataset of 128 samples. Additionally, as the reported thallium concentration is 0.01 µg/L greater than the screening value, it is considered inconsequential to overall potential risks. No maximum constituent concentrations exceeded either human health screening values or RBCs. Thus, no potential for risks to humans exposed to groundwater from the transition zone aquifer were identified. Crutchfield Branch – Surface Water One surface water sample location (SW-CB2) was evaluated in the Crutchfield Branch area (Figure 12-1). Data for the surface water sample location is included in Appendix B, Table 2. The 2016 risk assessment identified potential risks under a hypothetical recreational and subsistence fisher scenario exposed to cobalt in fish tissue modeled from surface water concentrations. The risks were likely overestimated because of very conservative assumptions in the exposure models. New surface water data indicate that cobalt concentrations have decreased since completing the 2016 risk assessment. Cobalt was not detected (<1 µg/L) April 7, 2017 in SW-CB2, indicating that potential for unacceptable health risks under the hypothetical fisher scenario are not likely when additional data are incorporated into the assessment. Detected concentrations of cobalt in surface water samples exceeded the 2L; however, 2L exceedances were also noted in upgradient reference location (SW-REF1). Except for one high turbidity sample, detected cobalt concentrations in surface water did not exceed the groundwater provisional background concentrations. Additionally, there is not evidence of cobalt in groundwater downgradient of the ash basin. These observations indicate that the presence and concentrations of cobalt in SW-CB2 is likely naturally occurring or from sources other than the ash basin. No evidence of risks to humans exposed to surface water in the downgradient Crutchfield Branch area was identified. “South Creek” – Surface Water One surface water sample location (S-06) is a reference location for the current CSA. Data for the surface water sample location are included in Appendix B, Table 2. No evidence of risks to humans exposed to surface water in the south creek area was identified. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 12-4 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx 12.2 Ecological Screening Summary Crutchfield Branch (off-site) – Surface Water One surface water sample location is used in the ecological assessment of Crutchfield Branch (SW-CB2). Data for the surface water sample are included in Appendix B, Table 2. The 2016 risk assessment resulted in no potential risks to wildlife from exposure to Crutchfield Branch surface water. Dissolved copper (0.45 micron filter) was detected in one of five surface water samples collected from location SW-CB2 (4.14 µg/L in April 2017). Total copper reported in this sample was below detection, or less than 1 µg/L, and turbidity in the sample was elevated at 25.1 NTU. The reported detection of dissolved copper is likely anomalous due to being greater than the total copper concentration and is not considered further with respect potential ecological risks. No potential unacceptable risks to wildlife exposed to surface water in Crutchfield Branch were identified. “South Creek” – Surface Water One surface water sample location (S-06) was used in the ecological assessment of the South Creek area. This location is a reference location for the current CSA. Data for the surface water sample location are included in Appendix B, Table 2. No potential risks to wildlife exposed to surface water in the south creek area were identified. 12.3 Private Well Receptor Assessment Update An independent study was conducted that evaluated 2015 groundwater data collected from 3 private drinking water wells within close proximity (<0.5 miles) of the Mayo Steam Electric Plant and 14 private drinking water wells within a 2 to 10 mile radius of the Mayo Steam Electric Plant (CAP 2, Section 5.7; Haley & Aldrich, 2015). Pertinent observations presented in the study included: Boron and arsenic were not detected in private wells sampled by NCDEQ; Calcium, sulfate, vanadium, and hexavalent chromium detected in the private well samples were less than their respective background threshold values; and Groundwater flow paths from the Mayo plant are away from residential areas. The Haley & Aldrich report concluded that the constituents detected in the private wells sampled by NCDEQ are consistent with regional background and do not indicate impact from constituents derived from coal ash. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 12-5 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx Recent (2017) results from off-site water supply wells indicate that constituent concentrations are less than 2L or less than PBTV for bedrock wells, with the exception of one manganese and one vanadium detection. Manganese was detected in samples collected from bedrock wells at 586 µg/L, compared to the bedrock PBTV of 544 µg/L, and vanadium was detected at 6.2 µg/L, compared to the bedrock PBTV of 5.52 µg/L. Based on these observations, there are no indications that potential for risks to off-site residences exposed to groundwater exist. 12.4 Risk Assessment Update Summary Based on review and analysis of groundwater and surface water data, there is no evidence of risks to humans and wildlife at the Mayo Site attributed to CCR constituent migration in groundwater from the ash basin. This update to the human health and ecological risk assessment supports a risk classification of “Low”. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 13-1 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx 13.0 GROUNDWATER MODELING RESULTS Groundwater flow, transport, and geochemical models are being developed to simulate movement of COIs through the subsurface to support the evaluation and design of remedial options at the Site. The models will provide insights into: 1. COI Mobility: Geochemical processes affecting precipitation, adsorption and desorption onto solids will be simulated based on analytical data and thermodynamic principles to predict partitioning and mobility in groundwater. 2. COI Movement: Simulations of the groundwater flow system will be combined with estimates of source concentrations, sorption, effective porosity, and dispersion to predict the paths and rates of constituent movement at the field scale. 3. Scenario Screening: The flow, transport, and geochemical models will be adjusted to simulate how various ash basin closure design options and groundwater remedial technologies will affect the short-term and long-term distribution of COIs. 4. Design: Model predictions will be used to help design basin closure and groundwater corrective action strategies in order to achieve compliance with PBTVs and/or 2L in a reasonable cost and time frame. The groundwater flow model linked with the transport model will be used to establish transport predictions that best represent observed conditions at the Site particularly for the constituents, such as boron, that tend to be negligibly affected by geochemical processes. The geochemical model information will provide insight into the complex processes that influence constituent mobility, which will be used to refine constituent sorption within the transport model. Once the flow, transport, and geochemical models for the Site accurately reproduce observed Site conditions, they can be used as predictive tools to evaluate the conditions that will result from various remedial options for basin closure (No Change, Cap-in-Place, or Ash Removal) and potential subsequent passive or active groundwater remedial technologies. The site-specific groundwater flow and transport models and the site-specific geochemical models are currently being updated for use in the CAP. The CAP will further discuss the purpose and scope of both the groundwater and geochemical models and will detail model development, calibration, assumptions and limitations. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 13-2 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx The CAP will also include a detailed remedial option evaluation based on observed conditions and the results of predictive modeling. The evaluation of the potential remedial options will include comparisons of predictive model results for long-term source concentration and plume migration trends toward potential receptors. The model predictions will be used in combination with other evaluation criteria to develop the optimal approach for basin closure and groundwater remediation. The following sections provide a brief summary of modeling efforts completed to date. 13.1 Summary of Fate and Transport Model Results A groundwater flow and transport model was developed to gain an understanding of COI migration after closure of the ash basin at the Mayo Plant. The initial groundwater model in the CAP, Part 1 (SynTerra, 2015b) included a calibrated steady-state flow model of June 2015 conditions, a calibrated historical transient model of constituent transport to June 2015 conditions, and three potential basin closure scenarios. Those basin closure simulation scenarios included: No change in Site conditions (basin remains open, as is) Cap-in-place Ash removal (excavation) The initial model used boron and arsenic as primary modeling constituents. As part of the CAP, Part 2 (SynTerra, 2016a) the model was revised to include manganese as a constituent for evaluation of future simulations. Additionally, the model predictive time was extended from 30 years to 100 years. The revised model in the CAP, Part 2 (SynTerra, 2016a) included a calibrated steady- state flow model of June 2015 conditions, a calibrated historical transient model of constituent transport to June 2015 conditions, and two potential basin closure scenarios. Those basin closure simulation scenarios included: No change in Site conditions (basin remains open, as is) Cap-in-place The flow and transport model is currently being updated as a part of the final CAP and will include: development of a calibrated steady-state flow model that includes data available through November 2017, development of a historical transient model of 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 13-3 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx constituent transport, and predictive simulations of basin closure plus groundwater corrective action scenarios. The updated fate and transport model will consider arsenic, boron, and possibly additional COIs that are hydraulically driven. Predictive simulations will have simulation times that continue until modeled COIs concentrations are below the 2L standard at the compliance boundary. The following sections provide a brief summary of the groundwater modeling that was presented in the CAP, Part 2 and a general outline for the updated modeling effort. The summary of the groundwater modeling presented in the CAP, Part 2 was compiled to address specific questions regarding model set-up and calibration. A complete updated groundwater flow and transport model report is being developed and will be submitted as part of the updated CAP. The model was developed using the MODFLOW-NWT version (Niswonger, Panday, & Motomu, 2011). This version provides improved numerical stability and accuracy for modeling problems within a variable water table. The improved numerical stability and accuracy can provide better estimates of the water table fluctuations that result from ash basin operating conditions and potential closure and groundwater corrective action activities. MT3DMS was used to simulate fate and transport of selected COIs. 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 the soil matrix adsorption of COIs. Flow Model Construction 13.1.1 The flow and transport model was built through a series of steps. The first step was to build a three-dimensional (3D) model of the Site hydrostratigraphy based on the SCM. The next steps were to determine the model dimensions and the construction of the numerical grid. The numerical grid was then populated with flow parameters, which were calibrated in the steady-state flow model. 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 time-dependent flow field for the constituent transport simulations. Generally, the model geometry will not be substantially modified for the updated model. Hydraulic parameters such as hydraulic conductivity values 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 13-4 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx may be adjusted within reasonable site-specific ranges to achieve hydraulic head calibration error below 10 percent. Flow Model Domain and Grid Layers The model has dimensions of approximately three miles-by-three miles, with the ash basin at the center of the model domain. The model domain was rotated 31˚ clockwise so that the model boundaries were parallel with the ash basin dam. The shortest distance between the ash basin and a model boundary is approximately one mile. The hydrostratigraphic model consists of six units: ash, saprolite, transition zone, upper bedrock (upper fractured rock), middle bedrock (middle fractured rock), and lower bedrock (lower rock). The units were determined by interpolating boring log data from historical data, the CSA, and the CAP reports. The hydrostratigraphic model was developed using “Solids” in GMS and was subdivided into five solids. A computational mesh (numerical grid) was then developed based on these solids: ash, saprolite, transition zone, fractured rock, and rock. The numerical grid consists of rectangular blocks arranged in columns, rows, and layers. There are 171 columns, 232 rows, and 15 layers. The maximum width of the columns and rows is 100 feet. The size of the grid blocks is approximately 50 feet by 50 feet in the vicinity of the ash basin. The horizontal dimension of some of the grid blocks is as small as 25 feet in the vicinity of the dams. The grid consists of 15 layers representing the six hydrostratigraphic units. It is expected that the updated model will use similar grid spacing. Hydrostratigraphic layer Grid layer Ash 1-4 Saprolite 5 Transition zone 6-7 Upper fractured rock 8-10 Middle fractured rock 11-12 Lower Rock 13-15 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 13-5 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx Flow Model Boundary Conditions The south, east, and north boundaries of the model were set to a specified head within the transition zone, which is a reasonable approximation of Site conditions. The specified head boundaries along the upland areas were terminated within a few hundred feet of the locations of streams or lakes that crossed those boundaries. Lakes are defined as specified heads. Streams are defined as drain-type boundaries. No-flow boundaries were set along groundwater divides. A no-flow boundary was also set beneath the transition zone to the base of the model. Sources and Sinks Water can enter the model or leave the model through the use of sources and sinks. MODFLOW uses point sources/sinks as well as areal sources/sinks. Point sources/sinks include rivers, wells, drains, and general head. Areal sources/sinks considered are limited to recharge. Source (Recharge) Model recharge sources include: Recharge that is applied to the ash basin (7.9 inches/year) and zero recharge applied to the inundated portion of the ash basin. (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 evapotranspiration 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.) Rainwater that infiltrates in the upland areas (6.5 inches/year) (Precipitation in developed areas of the Site (set to near zero; assumes most will run off). Large areas of ponded water, such as the ash basin and Mayo Lake, were represented as specific head boundaries and recharge was set to zero.) 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 13-6 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx Constant head boundaries (Lakes were represented as specified head boundaries with the head set to their stage. This includes Mayo Lake and the ash basin. The stage of the ash basin was set to 480 feet based on LIDAR data and a surveying point. The stage of Mayo Lake was 432 feet.) Model Sinks (Drains) Model sinks include: Streams within the model domain The former stream valley of Crutchfield Branch Engineered toe drains located at the base of the ash basin dam AOWs at the toe of the dam that feed smaller, unnamed tributaries to Crutchfield Branch Ash basin water diverted into a forebay, which is discharged through NPDES Outfall 002 into Mayo Lake. Water Supply Wells Approximately 21 domestic wells were previously identified within one-half mile of the Site (SynTerra, 2014a). The average daily use for domestic wells was set at a discharge of approximately 350 gallons per day (USEPA, 2017c). Hydraulic Conductivity The horizontal hydraulic conductivity and the horizontal-to-vertical hydraulic conductivity anisotropy ratio (anisotropy) are the main variable hydraulic parameters in the model. The distribution of those parameters is based primarily on the model hydrostratigraphy, with some local variations. The values can be adjusted during the calibration process to provide a best fit for observing water levels in wells. Initial estimates of parameters were based on literature values, results of slug and core testing, and simulations performed using a preliminary flow model. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 13-7 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx Streams and Lake Hydraulic Parameters Mayo Lake and the ash basin were represented as specified head boundaries. The stage of the ash basin was set to 480 feet based on LIDAR data from 2012 and surveying data. The stage of Mayo Lake was set to 432 feet and was also based on NCDOT LIDAR data from 2012. The outflow channel from the ash basin to the forebay was represented as specified head. Survey data of the outflow channel heads were collected and applied to the model that allows this engineered channel to exchange water with the groundwater system. Flow Model Calibration Targets The steady state flow model calibration data for June 2015 were presented in the CAP, Part 2. In the final CAP, calibration target data will be incorporated by taking the mean of the hydraulic head data for each well and applying a standard deviation to reflect the seasonal changes in the hydraulic heads. Hydraulic head data will include measurements until November 2017. Mass Balance The previous model had a mass balance error of well below 1%. The updated model will have a similar numerical accuracy. Flow Model Sensitivity Analysis A parameter sensitivity analysis for the preliminary calibrated model showed the highest degree of sensitivity to upland recharge and hydraulic conductivities (in the transition zone and saprolite stratigraphic units). The model was only weakly sensitive to the hydraulic conductivities of the ash, deep bedrock, and hydraulic conductivity of the dams and to the pumping rate of the domestic wells. Since no major elements within the model are to be changed, there is no need to perform additional sensitivity testing. Particle Tracking A primary concern is the potential impact to domestic and public wells from COIs migrating from the Site. The final calibrated groundwater flow model will be used to assess potential impacts by considering pumping from domestic and public wells, if any, within the model domain. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 13-8 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx Flow Model Assumptions and Limitations The groundwater model is currently being updated/refined and assumptions and limitations will be subject to change. Based on the preliminary modeling results, the assumptions and limitations included the following: The steady-state flow model was calibrated to hydraulic heads measured in monitoring wells in June 2015. The model was not calibrated to transient water levels over time, recharge, or stream flow. MODFLOW simulates flow through porous media. A single domain MODFLOW modeling approach for simulating flow in the primary porous groundwater zones and bedrock was used for contaminant transport. Flow in fractured bedrock is simulated using the equivalent porous media approximation. For the purposes of numerical modeling and comparing predictive scenarios, it was previously assumed that basin closure would be completed in 2015. A similar assumption will be used in the updated model. Predictive simulations were performed and steady-state flow conditions were assumed from the time that the ash basin was placed in service through the current time until the end of the predictive simulations (2045). The uncertainty in model parameters and predictions has not been quantified; therefore, the error in model predictions is not known. It was assumed the model results are suitable for a relative comparison of closure scenario options. Residential wells for which well construction records could not be obtained were assumed to be completed in the upper bedrock. Transport Model Construction 13.1.2 Modular 3-D Transport Multi-Species (MT3DMS) is being used to simulate constituent transport. MT3DMS simulates 3D advection and dispersion of the dissolved COIs, including the effects of retardation due to the soil matrix adsorption of COIs based on flow fields established by MODFLOW. The initial model used boron and arsenic as primary modeling constituents. The updated fate and transport modeling will focus on arsenic, boron, and possibly additional 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 13-9 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx COIs that are hydraulically driven. Other constituents will be considered in geochemical modeling. Transport Model Parameters The key transport model parameters (besides the flow field) are the constituent source concentrations 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. Transport Model Boundary Conditions In the current model, the transport model boundary conditions are “no flow” on the exterior edges of the model. 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 reach a drain or are removed by flow that enters a constant head boundary. In the current model, the concentrations from the June 2015 sampling event were set as boundary conditions within the ash basin. These values will be updated to use the concentration data up through the November 2017 sampling event. Transport Model Sources and Sinks Transport model sources include: The ash basin is considered the source of COIs in the model. The sources are simulated by applying a constant COI concentration within the cells of the ash basin and were applied to layers 1 through 4 which represent the ash. This allows infiltrating water to carry dissolved constituents from the ash pore water into the groundwater underneath the ash basin. Chemical analyses from four wells were used to characterize the distribution of COI concentration within the ash basin, and the source concentration is used as a calibration parameter in the transport simulations The concentration in the vicinity of ABMW-4 in Layer 5 to reflect the geometry of the ash basin. As the COIs migrate beneath and away from the coal ash, zones of soil and fractured rock may become impacted. These impacted zones can serve as secondary sources and are fully accounted for in the transport models. For simulations that involve ash excavation, the constant 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 13-10 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx concentration sources in the ash zones are removed, but the secondary sources in the impacted soil and fractured bedrock remain. The longevity of these secondary sources depends on the COI Kd and on the degree of flushing by infiltration and groundwater flow. Transport model sinks include: Lakes Streams Other AOW and engineered drains Transport Model Calibration Targets and Sensitivity The initial transport model calibration targets were COI concentrations measured in monitoring wells in June 2015. The updated model calibration targets will include COIs concentrations measured in monitoring wells in 2017. Constituents considered for the next fate and transport model will include boron, arsenic and possibly other COIs. COIs not amenable to simulation in the fate and transport model will be addressed in the geochemical model. Transport Simulation The updated model will be calibrated to include data through November 2017 and will extend until modeled COI concentrations are below the 2L standard at the compliance boundary. The following is a summary of the basin closure options modeled: No Action – Leave the ash basin as is to evaluate whether groundwater quality would be restored by natural attenuation under current conditions. Cap-in-Place – Grade the ash and place an engineered low permeability cover system to reduce infiltration of surface water. This scenario assumes that the ash under the cap will be dewatered. Ash Removal – Remove the ash from the basin. This scenario assumes that the ground surface would be restored to its initial grade (prior to construction of the ash basin). 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 13-11 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx The distribution of recharge, locations of drains, and distribution of material will be modified to represent the different basin closure options. The results of these simulations will be included as part of the updated CAP submittal. Summary of Flow and Transport Modeling Results To 13.1.3 Date The simulated June 2015 concentration distributions described in the CAP, Part 1 (SynTerra, 2015b) were used as initial conditions in a predictive simulation of future flow and transport at the Site and modeled arsenic and boron. Predictive simulations of future flow and transport for manganese (SynTerra, 2016a) were added and predictive simulations of future flow and transport for both boron and manganese under the “no action” and cap-in-place scenarios were run for a 100-year projection. No Action In the CAP, Part 1 report (SynTerra, 2015b) 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. The bottom of the transition zone and underlying rock are less than the 2L standard in 2045, according to the simulations. It should be noted the Kd value for arsenic was on the low side of the range of lab values and the simulation over-estimated several of the observed concentrations during calibration. The CAP, Part 1 (SynTerra, 2015b) indicates 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 leading edge of the boron plume is on the south side of Mayo Lake Road 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. The groundwater model as reported in the CAP, Part 2 (SynTerra, 2016a) indicates simulated 2115 boron concentrations delineated by the 2L contour are similar in 2045 simulations. The 2115 boron simulations within the transition zone and bedrock are slightly larger than the 2045 simulations. The 2115 boron plume simulations show the edge of the plume within the saprolite and transition zone is south of Mayo Lake Road. The simulated 2115 boron bedrock plume has two locations where the edge of the plume has migrated slightly north of Mayo Lake Road. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 13-12 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx The groundwater model as reported in the CAP, Part 2 (SynTerra, 2016a) indicates that the observed concentrations of manganese in the ash basin would result in a plume of manganese extending along Crutchfield Branch. This plume would occur in a setting where manganese was common as a natural constituent of groundwater. The simulations indicate that approximately ¾ of the wells where manganese exceeded the 2L standard represent naturally occurring manganese that is not included in the model. The other ¼ of the affected wells are located in a region influenced by the manganese plume in the simulations. 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. 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. Cap-In-Place In the CAP, Part 1 report (SynTerra, 2015b), simulated arsenic concentrations in the saprolite beneath the ash basin slightly increased from 2015 to 2045 and are less than the No Action scenario. According to the simulations, arsenic is below the 2L standard in the transition zone and fractured bedrock. It should be noted the Kd value for arsenic was on the low side of the range of lab values and the simulation over-estimated several of the observed concentrations during calibration. Under a cap-in-place (engineered capping system) closure scenario, the model results indicate a stabilization of boron plume geometry and a reduction of boron concentrations in the surficial zone by 2030 (SynTerra, 2015b). Likewise, for the transition zone, plume extent and concentrations decrease by 2030 and remain stable through 2045. For the fractured bedrock, the plume “footprint” begins to retreat south away from the compliance boundary by 2030 and continues to retreat south by 2045. Beginning with the 2045 simulation and continuing until 2115 (100 years), the simulations indicate continued boron concentration reductions and additional plume retreat by several hundred feet by 2115 for all units. For all three hydrogeologic units, no 2L exceedances of boron beyond the 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 13-13 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx compliance boundary are indicated by the simulation results. Results in the February 2016 updated fate and transport model indicate that the leading edge of the boron plume recedes from 2045 to 2115 within the saprolite, transition zone, and fractured bedrock layers up to 300 feet within the vicinity of Mayo Lake Road. Model results indicate that the leading edge of the highest manganese concentration having a source within the ash basin has begun to recede approximately 30 years after basin closure (2045). More significantly, after 100 years (2115), the manganese plume in exceedance of the 2L is 100 feet within the compliance boundary. The results of the CAP, Part 2 modeling did not substantially alter the conclusions presented in CAP, Part 1. 13.2 Summary of Geochemical Model Results The Mayo Site geochemical model investigates how variations in geochemical parameters affect movement of constituents through the subsurface. The geochemical site conceptual model (SCM) will be updated as additional data and information associated with Site constituents, conditions, or processes are developed. The geochemical modeling approach presented in the following sections was developed using laboratory analytical procedures and computer simulations to understand the geochemical conditions and controls on groundwater concentrations at the Mayo Plant in order to predict how remedial action and/or natural attenuation may occur at the Site and avoid unwanted side effects. The final geochemical model will be presented in the updated CAP. Model Construction 13.2.1 The geochemical model in the CAP, Part 2 (SynTerra, 2016a) included: EH-pH (Pourbaix) diagrams showing potential stable chemical phases of the aqueous electrochemical system, calibrated to encompass conditions at the Site; Sorption model where the aqueous speciation and surface complexation are modeled using the USGS geochemical modeling program pH, redox, equilibrium model written in C language (PHREEQC); 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 13-14 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx Simulations of the anticipated geochemical speciation that would occur for each COI in the presence of adsorption to soils and in response to changes in redox potential (EH)and pH; and Attenuation calculations where the potential capacity of aquifer solids to sequester constituents of interest were estimated. Laboratory Determination of Distribution Coefficient SynTerra retained researchers from the UNCC to determine site-specific distribution coefficients (Kd) for the primary hydrostratigraphic units. The UNCC Soil Sorption Evaluation and Addendum to the UNCC Soil Sorption Evaluation reports are provided in Appendix C. Selected soil samples were analyzed using batch and column experiments to determine Kd values for COIs (Table 13-1). In addition to these analyses, metal oxy-hydroxide phases of iron (HFO), manganese (HMO), and aluminum (HAO) in soils were measured. HFO, HMO, and HAO are considered to be the most important surface reactive phases for cationic and anionic constituents in many subsurface environments (Ford, W., & Puls, 2007). Quantities of these phases in soil can thus be considered a proxy for the presence of ferrihydrite (HFO) and gibbsite (HAO) which can be used to model COI sorption capacity for a given soil (Dzombak & Morel, 1990); (Karamalidis & Dzombak, 2010). Geochemical Model Construction To examine the sorption behavior of multiple ions of interest in the subsurface environment surrounding coal-fired power plants, a combined aqueous speciation and surface complexation model was developed using the USGS geochemical modeling program PHREEQC. Equilibrium constants for aqueous speciation reactions were taken from the USGS WATEQ4F database. This database contained the reactions for most elements of interest except for antimony, chromium, cobalt, and vanadium. Constants for aqueous reactions and mineral formation for these elements were taken from the MINTEQ v4 database which is also issued with PHREEQC. The constants were checked to provide a self-consistent incorporation into the revised database. The source of the MINTEQ v4 database is primarily the well-known NIST 46 database (Martell & Smith, 2001). Sorption reactions were modeled using a diffuse double layer surface complexation model. For self-consistency in the sorption model, a single database of constants was used as opposed to searching out individual constants from literature. The diffuse double layer model describing ion sorption to HFO 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 13-15 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx and HAO was selected for this effort (Dzombak & Morel, 1990); (Karamalidis & Dzombak, 2010). Geochemical Controls on COI As described in previous geochemical model reports (SynTerra, 2015b; SynTerra, 2016a), pH, EH, and solubility are the primary geochemical parameters affecting constituent mobility. In the updated geochemical model planned to be submitted in 2018 as part of the CAP, hydraulically significant flow transects will be used to evaluate the conceptual model of COI mobility in the subsurface. Trends in the concentrations of each COI along the transect will be compared with the model output to verify that the conceptual and qualitative models can predict COI behavior. The model will then be used to evaluate the potential impacts of remediation activities. The model will relate the COI concentrations observed in groundwater along flow transects to key geochemical parameters influencing constituent mobility (i.e., EH, pH, and saturation/solubility controls). Geochemical Model Assumptions Several key assumptions will be applied to the planned geochemical modeling effort: 1) The thermochemical sorption constant reactions describe ion sorption to ferrihydrite and gibbsite (HFO and HAO). 2) The model will use the same or more conservative site density assumptions as those used by Dzombak and Morel (1990) and Karamalidis and Dzombak (2010) to constrain the surface sites. HAO and HFO (i.e., gibbsite and ferrihydrite) are used as the primary reactive minerals due to the availability of surface complexation reactions. Differences between the sorption behaviors at a specific Site will be primarily due to 1) differences in the pH, EH, and ion concentrations, and 2) differences in the extractable iron and aluminum concentrations from site-specific solids. Additional reactive minerals will be incorporated into the model as needed on a site-specific basis. Updated Geochemical Model Development The updated geochemical site investigation to accompany the CAP will develop parameters for each aquifer or geologically derived flow zone (geozone) by considering the bulk densities, porosities, and hydraulic gradients used in the fate and transport model. Additionally, the potential effects on aquifer 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 13-16 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx characteristics resulting from the construction of a future retention basin and flue-gas desulfurization pond will be evaluated. These parameters are used to constrain the sorption site concentrations in the model input and will be incorporated in the 1-D ADVECTION model to accompany the capacity simulations. The objective of these capacity simulations is to determine the mass balance on iron and aluminum sorption sites when simulating flow through a fixed region. Input and initial iron and aluminum concentrations will be fixed based on site-specific data. Thus, the model will be able to simulate the stability of the HFO and HAO phases assumed to control constituent sorption. The final geochemical model report will include a site specific discussion of: The model description, The purpose of the geochemical model, Modeling results with comparison to observed conditions, COI sensitivity to pH, EH, iron/aluminum oxide content, and Model limitations. The updated geochemical modeling will also present multiple methods of determining constituent mobility at the Site. Aqueous speciation, surface complexation, and solubility controls will be presented in the revised report. These processes will be modeled using: Pourbaix diagrams created with the Geochemist Workbench v10 software using site-specific minimum and maximum constituent concentrations. PHREEQC’s combined aqueous speciation and surface complexation model and the 1-D ADVECTION function to gain a comprehensive understanding of current geochemical controls on the system and evaluate how potential changes in the geochemical system might affect constituent mobility in the future. Summary of Geochemical Model Results To Date 13.2.2 The geochemical model considers changes in oxidation state for all redox active constituents of interest (arsenic, chromium, cobalt, iron, manganese, selenium, sulphur, and vanadium) and changes in chemical speciation for all constituents. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 13-17 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx PHREEQC model predicts: Arsenic - As(V) as the dominant oxidation state of arsenic. As(III) is the dominant species measured in ground waters. This is due to the stronger sorption of As(V). This results in the relatively lower Kd values predicted for arsenic in the model. Boron - Boron exists only in the B(III) oxidation state. PHREEQC predicted Kd values for boron are low. Predicted values are slightly lower, but generally consistent, with the values chosen for reactive transport modeling and those measured in batch laboratory experiments. Chromium - Cr(III) is the dominant oxidation state which. Cr(III) readily sorbs to mineral surfaces as the pH increases. Cr (VI) however, is weakly sorbing and decreases sorption as the pH increases. Kd for Cr (III) is relatively high while the Kd for Cr (VI) is relatively low. Manganese - Manganese is predominantly present as Mn2+. Sorption of Mn(II) is generally weak and yields low Kd values. Manganese bearing soil minerals could occur given sufficiently high manganese concentrations and high pH/ EH conditions which may play a role on controlling the movement of manganese in the subsurface. Cobalt - The dominant cobalt species predicted by the PHREEQC model is Co2+ redox potential exhibits relatively little influence on Co2+. Overall, cobalt is expected to exhibit minimal transport in these systems (high Kd) relative to more mobile species. Selenium – The geochemical behavior of selenium is highly dependent on the EH of the groundwater. The PHREEQC model predictions show Se(IV) as the dominant species under approximately neutral pH conditions but the fraction of Se(VI) increases with increasing pH and EH. Overall the range of Kd values predicted by PHREEQC agrees with the values determined experimentally from batch sorption tests. Vanadium - Vanadium can exist in multiple oxidation states including V(III), V(IV), and V(V) under the groundwater conditions at the Site. The majority of vanadium is expected to exist as pentavalent V(V) which exhibits moderate sorption. Predicted Kd values are highly dependent on the pH of the groundwater. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 13-18 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx 13.3 Summary of Groundwater to Surface Water Evaluation For the scope of this CSA, it is only appropriate to compare named surface waters to 2B because AOWs, wastewater, and wastewater conveyances (effluent channels) are evaluated and governed wholly separate in accordance with the NPDES Program administered by NCDEQ DWR. This process is on-going in a parallel effort to the CSA and subject to change. Prior to construction of the ash basin, Crutchfield Branch was a perennial stream that originated about 1,000 feet southwest of the current ash basin footprint. The ash basin now encapsulates the former headwaters of Crutchfield Branch and two smaller, intermittent streams that flowed into Crutchfield Branch. Groundwater underlying the ash basin flows north-northeast and beneath and through the base of the ash basin dam and into the former Crutchfield Branch stream valley. Two discharges from engineered toe drains emerge from the area around the ash basin dam and contribute to the volume of water flowing downstream of the dam. Surface water flows northeast beneath Mayo Lake Road, onto privately-owned property where it merges with another small tributary that originates on the Mayo Site, and then flows north into Virginia. Consistent with LeGrand’s model for groundwater flow and discharge in the Piedmont (LeGrand, 2004) the Crutchfield Branch stream valley and its tributaries are groundwater discharge zones downgradient of the ash basin. Groundwater data collected to date and groundwater flow and transport modeling performed in CAP, Part 2 indicate that surface water in the area downgradient of the ash basin is influenced by groundwater discharge. As described in Section 9.0, boron concentrations are greatest near and proximate to the engineered toe drains and decrease from the Mayo Lake Road to the state line. Manganese concentrations are similarly consistent immediately downstream of the ash basin and begin to decline downstream; however, at the most downstream surface water location (SW-CB2), manganese concentrations increase. Groundwater modeling shows that surface water in the area downgradient of the ash basin is influenced by groundwater discharge. The CAP will include updated modeling and use those results, in conjunction with existing data as needed, to determine if the proposed corrective action will result in exceedance of surface water quality standards 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 13-19 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx in Crutchfield Branch. Closure and bulk dewatering of the ash basin is expected to reduce groundwater contribution to surface water. Basin closure modeling will be used to predict groundwater to surface water contribution. The updated CAP will evaluate if corrective action following basin closure will be necessary to mitigate a potential exceedance of surface water quality standards. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 14-1 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx 14.0 SITE ASSESSMENT RESULTS 14.1 Nature and Extent of Contamination The site assessment described in the CSA presents the results of investigations required by CAMA and 2L regulations. The ash basin pore water was determined to be a source of impact to groundwater. The site assessment investigated the Site hydrogeology, determined the direction of groundwater flow from the ash basin, and determined the horizontal and vertical extent of impacts to groundwater and soil sufficient to proceed with preparation of a CAP. Constituents of Interest COIs in groundwater identified as being associated with the Mayo Plant ash basin include arsenic, barium, boron, chromium, hexavalent chromium, cobalt, iron, manganese, molybdenum, pH, strontium, sulfate, TDS, and vanadium. Groundwater COIs migrate laterally and vertically into and through surficial regolith, the regolith/bedrock transition zone, and shallow bedrock. The surficial zone at the Mayo Site is generally thin and unsaturated. The transition zone, where saturated, and shallow bedrock generally contain the first occurrence of groundwater. Constituent migration in groundwater occurs at variable rates depending on a number of physical and chemical conditions and properties (e.g., constituent sorption properties, redox state, pH, hydraulic conductivity, etc.). Some COIs, such as boron, readily solubilize and migrate with minimal retention. In contrast, some COIs such as arsenic readily adsorb to aquifer materials, do not readily solubilize, and thus are relatively immobile. Hydrogeologic Conditions Site hydrogeologic conditions were evaluated by installing and sampling groundwater monitoring wells and piezometers; conducting in-situ hydraulic tests; sampling soil for physical and chemical testing; and sampling surface water, AOW, and sediment samples. Monitoring wells were completed in each hydrostratigraphic unit. The groundwater flow system serves to store and provide a means for groundwater movement. The porosity of the regolith is largely controlled by pore space (primary porosity); whereas, in bedrock, the effective porosity is largely secondary and controlled by the number, size, and interconnection of fractures. The nature of groundwater flow across the Site is based on the character and configuration of the ash basin relative to specific Site features such as man-made and natural drainage features, engineered drains, streams, and lakes; hydraulic boundary conditions; and subsurface media properties. The majority of groundwater flow across the Site appears to flow through the transition zone and upper bedrock. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 14-2 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx Four hydrostratigraphic units were identified at the Mayo Plant and were evaluated during the CSA. A detailed description of each unit is provided in Section 6.2.2. Ash – The ash pore water unit consists of saturated ash material. Shallow/Surficial – The shallow/surficial unit consists of soil, saprolite, and alluvial material that overlie the transition zone with bedrock. Transition Zone – The transition zone flow unit lies directly above competent bedrock and is a zone of partially weathered bedrock. Fractured Bedrock – The majority of water-producing fracture zones were found within 50 - 75 feet of the top of competent bedrock. The Mayo Plant ash basin occupies the former stream valley of Crutchfield Branch. The basin acts as a bowl-like feature toward which groundwater flows from the northwest, west, south, and east. Groundwater flows north-northeast from the ash basin into the small valley formed by Crutchfield Branch. Groundwater flows from the highest topographic portion of the Site (near the Plant entrance road) to the north and northeast. The ash basin was formed when the Crutchfield Branch valley was dammed. The flow of ponded water within the ash basin is controlled laterally by groundwater flow that enters the basin from the east, south, and west and is controlled downgradient (north-northeast) by the ash basin dam and the NPDES outfall/discharge. The head created by the ash pore water creates a slight mounding effect that influences the groundwater flow direction in the immediate vicinity of the ash basin. East of the ash basin, there is a groundwater divide that separates the Crutchfield Branch flow regime from the Mayo Lake flow regime. There are few substantive differences in water level among wells completed in the different flow zones across the Site (shallow/surficial, transition zone, bedrock), and lateral groundwater movement predominates over vertical movement. The vertical gradients are near equilibrium across the Site, indicating that there is no distinct horizontal confining layer beneath the Mayo Plant. The horizontal gradients, hydraulic conductivity, and AOW velocities indicate that most of the groundwater transport occurs through the transition zone and bedrock, as most of the regolith encountered is largely unsaturated. Groundwater flow directions and the overall morphology of the potentiometric surface vary little from “dry” to “wet” seasons. Water levels do fluctuate up and down with significantly increased or decreased precipitation, but the overall groundwater flow 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 14-3 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx direction does not change due to seasonal changes in precipitation. Horizontal gradients along the southern portion of the Site range from 0.017 feet/feet to 0.018 feet/feet. Horizontal gradients along the northern end of the Site range from 0.03 feet/feet to 0.034 feet/feet. The hydraulic gradient in the northern portion of the Site is influenced by the higher relief. Generally, upward vertical gradients predominate on the west side of the Site, including the ash basin area, while downward (recharge) gradients are more prevalent in the northeast portion of the property. Upward vertical gradients from bedrock, as groundwater from the west, south, and east recharge the groundwater beneath the basin into the former Crutchfield Branch valley, reduces the potential for downward migration of COIs into bedrock. Horizontal and Vertical Extent of Impact The groundwater plume is defined as locations (in three-dimensional space) where groundwater quality is impacted by the ash basin. Naturally occurring groundwater contains varying concentrations of a number of constituents (e.g., alkalinity, aluminum, magnesium, sodium, zinc, etc.). Sporadic and low-concentration exceedances of these constituents in the groundwater data do not necessarily demonstrate distribution of groundwater that has likely been impacted by the ash basin. The leading edge of the plume, the furthest downgradient edge, is represented by groundwater concentrations greater than PBTVs in the wells in each flow unit. Boron is the primary CCR-derived constituent in groundwater at Mayo and is detected at concentrations greater than the PBTV and 2L standard beneath and downgradient (north-northeast) of the ash basin. Boron is not detected in background groundwater. Boron, in its most common forms, is soluble in water, and boron has a very low Kd value, making the constituent highly mobile in groundwater. Therefore, the presence/absence of boron in groundwater provides a close approximation of the distribution of CCR-impacted groundwater. The detection of boron at concentrations in groundwater greater than the PBTV best represents the leading edge of the CCR- derived plume moving downgradient from the source area (ash basin). At Mayo, boron is detected at concentrations greater than the 2L standard beneath and downgradient (north-northeast) of the ash basin. Boron is not detected in background groundwater. At Mayo, manganese and strontium detections in groundwater also indicate impact. The area farthest downgradient at which boron, manganese, and strontium are detected at a concentration greater than applicable PBTVs is interpreted as the leading edge of the CCR-derived plume moving downgradient from the source area. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 14-4 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx In the bedrock flow unit, boron is detected (does not exceed 2L) in downgradient well CW-2D (located on Mayo property, near the northern property line). In bedrock well MW-16BR, located approximately 1,000 feet downgradient of CW-2D and the property line, boron is not detected. The leading edge of the bedrock plume is interpreted to be at/near the northern property line. Similarly, in the transition zone unit, groundwater from monitoring well CW-2 contains detectable concentrations of boron that exceed 2L, while MW-16D, located about 1,000 feet downgradient, boron is not detected. Surficial wells MW-16S and MW-3 are screened in Crutchfield Branch alluvium. MW-3 (located on Mayo property several hundred feet south of the property line between the property line and the ash basin dam) contains detectable concentrations of boron that exceed 2L, while at MW-16S (located downgradient approximately 1,500 feet from MW-3), boron is detected at a lower concentration (does not exceed 2L). The surficial and transition zone flow units at Mayo – beneath and downgradient of the ash basin – are impacted by CCR-derived constituents; however, these units are not vertically extensive. Impact to the bedrock flow unit is confined, approximately, to the top 50 – 75 feet of fractured bedrock. The vertical extent of the plume is represented by groundwater concentrations in bedrock wells beneath and downgradient of the ash basin. ABMW-2BRL, drilled to a depth of 180 feet bgs, contains no boron or manganese concentrations above 2L or the PBTV. ABMW-2BR and ABMW-4BR are other shallower bedrock wells beneath the ash basin, and groundwater from these wells has the same absence of CCR-related constituents at concentrations that exceed 2L/IMAC or PBTVs. As groundwater under the ash basin flows northeast toward the ash basin dam, the hydraulic impact of the ash basin dam and the hydraulic head exerted by the ash basin water forces groundwater downward into the bedrock which increases hydraulic pressure in the bedrock aquifer. Wells completed in surficial, transition zone, and bedrock proximate to the north side of the ash basin dam are impacted by COIs. As groundwater and the plume migrate in the downgradient direction, unimpacted groundwater enters the system from upgradient recharge areas to the west and east, mitigating the concentration of some COIs (e.g., boron). 14.2 Maximum COI Concentrations Changes in COI concentrations over time are included as time-series graphs (Figures 14-1 through Figure 14-39). The maximum historical detected COI concentrations in groundwater for ash pore water or wells directly beneath the ash basin and non-ash basin groundwater are included below. Also listed is the range of PBTVs for the surficial, transition zone, and bedrock flow units: 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 14-5 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx Arsenic – Ash Basin: 1,020 µg/L (ABMW-2); Outside Basin: 2.05 µg/L (MW- 16BR); PBTV range: Not Detected in background samples for all three units. Barium – Ash Basin: 1,280 µg/L (ABMW-1); Outside Basin: 178 µg/L (CW-2); PBTV range: 19 µg/L – 97 µg/L. Boron - Ash Basin: 9,200 µg/L (ABMW-2); Outside Basin: 1,050 µg/L (MW-3); PBTV range: Not Detected in background samples for all three units. Chromium – Ash Basin: 1.39 µg/L (ABMW-4D); Outside Basin: 53 µg/L (CW-1); PBTV range: 3.23 µg/L – 7 µg/L. Chromium (hexavalent) – Ash Basin: 6 µg/L (ABMW-3); Outside Basin: 19.4 µg/L (MW-11BR); PBTV range: 0.088 µg/L – 1.26 µg/L. The Ash Basin value is questionable because the highest total chromium value is less than 6 µg/L. Cobalt – Ash Basin: 821 µg/L (ABMW-1); Outside Basin: 10.3 µg/L (MW-16S); PBTV range: Not detected (transition zone) – 1.19 µg/L. Iron – Ash Basin: 72,100 (ABMW-4D); Outside Basin: 11,700 µg/L (MW-5BR); PBTV range: 385 µg/L – 2,550 µg/L. Manganese – Ash Basin: 6,960 (ABMW-4D); Outside Basin: 2,680 µg/L (MW- 8BR); PBTV range: 253 µg/L – 544 µg/L. Molybednum – Ash Basin: 1,880 µg/L (ABMW-2); Outside Basin: 60.4 µg/L (MW-16BR); PBTV range: Not detected (transition zone) – 13.1 µg/L. pH - Ash Basin: 5.2 (ABMW-3S) – 9.6 (ABMW-2); Outside Basin: 4.9 (MW-4) – 9.3 (MW-16BR); PBTV range: 4.9 – 7.3. Strontium - Ash Basin: 4,640 µg/L (ABMW-1); Outside Basin: 1,100 µg/L (MW- 7BR); PBTV range: 25 µg/L – 418 µg/L. Sulfate – Ash Basin: 80 mg/L (ABMW-4BR); Outside Basin: 90 mg/L (MW-10BR); PBTV range: 1.6 mg/L – 18 mg/L. TDS – Ash Basin: 700 mg/L (ABMW-4D); Outside Basin: 810 mg/L (MW-19BR); PBTV range: 85 mg/L – 430 mg/L. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 14-6 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx Vanadium – Ash Basin: 314 µg/L (ABMW-1): Outside Basin: 5.9 µg/L (MW-12S)’ PBTV range: 0.974 µg/L – 5.88 µg/L. 14.3 Contaminant Migration and Potentially Affected Receptors Contaminant Migration The groundwater flow system at the Site serves both to store and provide a means for groundwater movement. The porosity of the regolith is largely controlled by pore space (primary porosity), whereas in bedrock porosity is largely controlled by the number, size and interconnection of fractures. As a result, the effective porosity in the regolith is normally greater than in the bedrock and thus the quantity of groundwater flow will be greater in the regolith. At the Mayo Plant, saturated regolith was observed in only a few wells, and the regolith is the least transmissive of the flow zones. The majority of groundwater across the Site appears to flow through the transition zone and bedrock. The pore water in the ash basin is the source of constituents detected above PBTVs or 2L in groundwater samples in the vicinity of the ash basin. Pore water analytical results are compared to 2L and/or IMAC for reference purposes only. The ash basin is a permitted wastewater system; therefore, comparison of pore water within the wastewater treatment residuals (ash) to 2B or 2L/IMAC is not required. Gradients measured within the ash basin support the interpretation that ash pore water mixes with shallow/surficial groundwater and migrates downward into the transition zone. Continued vertical migration of groundwater downgradient of the ash basin is also evidenced by detected constituent concentrations. Ash basin constituents become dissolved in groundwater that flows in response to hydraulic gradients. Groundwater migrates under diffuse flow conditions in the surficial aquifer in the direction of the prevailing gradient. As constituents enter the transition zone and fractured bedrock flow systems, the rate of constituent transport has the potential to increase. Groundwater flow is the primary mechanism for migration of constituents to the environment. At Mayo, groundwater movement in the bedrock flow zone is due primarily to secondary porosity represented by fractures in the bedrock. Primary (matrix) porosity is negligible; therefore, it is not technically appropriate to calculate groundwater velocity using effective porosity values. Bedrock fractures encountered at Mayo tend to be isolated with low interconnectivity. Further, hydraulic conductivity values measure the fractures immediately adjacent to a well screen, not across the distance between two bedrock wells. Groundwater flow in bedrock fractures is anisotropic and difficult to 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 14-7 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx predict, and velocities change as groundwater moves between factures of varying orientations, gradients, pressure, and size. The hydrogeologic characteristics of the ash basin environment are the primary control mechanisms on groundwater flow and constituent transport. The basin acts as a bowl- like feature toward which groundwater flows from all directions except from the north. The stream valley in which the ash basin was constructed is a distinct slope-aquifer system in which flow of groundwater into the ash basin and out of the ash basin is restricted to the local flow regime. Groundwater and surface water flow from the ash basin is funneled into the small valley formed by Crutchfield Branch. Boron, strontium, and manganese are present in groundwater downgradient of the ash basin on the Site near or at the compliance boundary in concentrations that exceed the 2L or PBTV. Figures 14-40 to 14-54 graphically depict the most recent available (March-April 2017) valid COI groundwater analytical concentrations for monitoring wells. Figures 14-55 to 14-69 show the most recent COI surface water and AOW concentrations. The figures are colored-coded to visually depict whether analytical concentrations seem to be increasing, decreasing, stable, or a trend could not be determined. Recent concentrations of COIs in groundwater, surface water, and AOWs are provided on Figures 14-70 and 14-71. Recent concentrations of COIs in solid media, as well as available geochemical properties of soils, are provided on Figure 14-72. Potentially Affected Receptors A baseline human health and ecological risk assessment was performed in 2016 as a component of the CAP, Part 2 (SynTerra, 2016a), concluding that no unacceptable risks to humans resulted from hypothetical exposure to constituents detected in the ash basin area. Based on review and analysis of groundwater and surface water data collected since completing the human health and ecological risk assessment in 2016, there is no evidence of potential risks to humans and wildlife at the Mayo Site. Water Supply Wells Results from private water supply wells did not indicate human health risks to off-site residents potentially exposed to groundwater associated with the ash basin. In addition, no public or private drinking water wells, supply wells, or wellhead protection areas were found to be located downgradient of the ash basin. Samples were collected from 11 private water supply wells located upgradient of the Mayo Plant ash basin to the northwest along US Highway 501, northwest in Virginia along US Highway 501, and south of the Plant around Mullins Lane. NCDEQ 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 14-8 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx coordinated sampling of three private water supply wells in 2015. Duke Energy collected samples from eight additional private water supply wells in 2017. A review of the analytical data for the private water supply wells indicated several constituents were detected above 2L or IMACs, including pH (two wells), iron (three wells), lead (one well), manganese (three wells), and vanadium (eight wells). In addition, concentrations of other COIs exceeded the respective PBTVs for a number of these private water supply wells. However, these data (2L, IMAC and PBTV exceedances) should be interpreted with caution for the reasons described below: PBTVs were developed using groundwater data from a set of three background wells all located on the Site. These wells are located within about 0.1 square miles of one another. The geochemical data from these wells may not be representative across the broader area encompassed by the 11 private water supply wells (spread across approximately 1.5 square miles). There is very limited information available about the sampled wells. Well construction equipment such as pipes, pumps, and fittings may influence water quality. A numerical capture zone analysis for the Mayo Site was conducted to evaluate potential impact of upgradient water supply pumping wells. The analysis indicated that capture zones from wells located to the northwest and southeast of the Mayo Plant are limited to the immediate vicinity of the well head and do not extend toward the ash basin. None of the particle tracks originating in the ash basin moved into the well capture zones. Further, groundwater flow direction relative to the ash basin has been demonstrated to be to the north-northeast. The private wells around the Mayo Plant are all located upgradient and away from the direction of groundwater flow. The geochemical signature of groundwater from the supply wells was compared with the signature of groundwater from the source area using Piper diagrams. The geochemical nature of groundwater from the sampled supply wells is very different from ash pore water and from groundwater beneath the basin. It is concluded that there is no impact to the supply wells that are located upgradient from the ash basin/Mayo Plant. The land directly downgradient of the ash basin and the Duke property line is undeveloped. Therefore, no water supply wells are located north and downgradient of the groundwater plume within the survey radius. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 14-9 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx No surface water intakes, other than the intake used to pump cooling tower make-up water from Mayo Lake for the Mayo Plant operations, are located in the vicinity of Mayo Plant either in another part of Mayo Lake or in Crutchfield Branch. Surface Water Three surface water samples were collected and analyzed, including one downstream of the ash basin and two reference, upstream locations. Surface water sample data indicates that only turbidity and DO concentrations exceeded a 2B criteria. Boron was detected in the downstream location (SW-CB2) but not in the reference locations. Strontium was detected in both SW-CB2 and in reference location SW-REF1; however, the detected concentrations in SW-CB2 were generally two to three times higher than in SW-REF1. Other constituents are higher in SW-CB2 surface water than in the reference locations (e.g., TDS, cobalt, iron, manganese); although, not several times higher. The majority of the flow in the former Crutchfield Branch stream valley is associated with engineered drainage from the ash basin immediately below the dam and other natural seepage of ash basin water which is evaluated and governed wholly separate in accordance with the NPDES Program administered by NCDEQ DWR. The groundwater in the area in deeper flow zones near the receiving stream generally contains constituent concentrations less than those of the receiving waters. Additional surface water sampling will be completed and an evaluation of potential impacts of groundwater on surface water will be presented in the CAP. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 15-1 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx 15.0 CONCLUSIONS AND RECOMMENDATIONS A discussion of preliminary corrective action alternatives that may be appropriate to consider during the updated CAP development are presented in this section. 15.1 Overview of Site Conditions at Specific Source Areas CCR material and pore water in the ash basin were determined to be a source of impact to groundwater. Boron is the primary constituent detected in groundwater at concentrations greater than background and the 2L near or beyond the compliance boundary; however, manganese and strontium also appear to be indicators of impact to groundwater. The interpreted extent of boron concentrations greater than the 2L standard is near the compliance boundary in the surficial and transition flow zones. The boron concentration is less than 2L standard in the bedrock flow unit near the compliance boundary. The interpreted extent of manganese concentrations greater than the PBTV and 2L standard is beyond the compliance boundary in the surficial and bedrock flow zones; however, the manganese concentration is less than the PBTV within the transition zone at the compliance boundary. The interpreted extent of strontium concentrations greater than the PBTV extends beyond the compliance boundary only within the surficial flow zone. 15.2 Revised Site Conceptual Model Site Conceptual Models (SCMs) are developed to be a representation of what is known or suspected about a site with respect to contamination sources, release mechanisms, transport, and fate of those contaminants. SCMs can be a written and/or be a graphic presentation of site conditions to reflect the current understanding of the site, identify data gaps, and be updated as new information is collected. SCMs can be used to develop an understanding of the different aspects of site conditions, such as a hydrogeologic conceptual site model to help understand the site hydrogeologic conditions affecting groundwater. SCMs can also be used in a risk assessment to understand contaminant migration and pathways to receptors. In the initial Site conceptual hydrogeologic model presented in the GAP (SynTerra, 2014c), the geological and hydrogeological features influencing the movement, chemical, and physical characteristics of contaminants were related to the Piedmont hydrogeologic system present at the Site. A preliminary SCM was developed from data generated during previous assessments, existing groundwater monitoring data, and CSA activities. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 15-2 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx The Mayo Plant has a single ash basin which contains ash generated from the Plant’s historic coal combustion. The ash basin is approximately 140 acres in size and is constructed with an earthen dike. CCR was managed at the Plant’s on-site ash basin and transported via wet sluicing until 2013. The 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 and eventually discharged through NPDES Outfall 002 into Mayo Lake. Borings installed in the ash basin encountered ash from 13.5 feet to 66.1 feet in thickness. Roughly 40 percent of the ash basin is covered with standing water. Assessment findings determined that CCR accumulated in the ash basin is the primary source of impact to groundwater. As previously discussed, residual concentrations of some COIs in soil beneath the ash basin may indicate limited impact to soil beneath the ash basin. The Mayo Plant ash basin occupies the former stream valley of Crutchfield Branch. The basin acts as an elongated bowl-like feature toward which groundwater flows from the northwest, west, south, and east. A small topographic high is present along the eastern side of the ash basin, and groundwater is somewhat radial away from this feature. Groundwater flow east of the railroad line, which is constructed along a natural ridge, is to the east and toward Mayo Lake. Groundwater flows north-northeast from the ash basin into the small valley formed by Crutchfield Branch. Crutchfield Branch flows north off of the Site into Virginia. Site-specific groundwater COIs were developed by evaluating groundwater sampling results with respect to 2L/IMAC and PBTVs, and additional regulatory input/requirements. The distribution of constituents in relation to the ash basin, co- occurrence with CCR indicator constituents such as boron, and likely migration directions based on groundwater flow direction were considered in determination of groundwater COIs. Wells monitoring the surficial, transition zone, and bedrock flow units were installed beneath the ash basin. Wells completed in the saprolite or transition zone beneath the ash basin have PBTV and 2L exceedances for arsenic, barium, boron, cobalt, iron, manganese, molybdenum, strontium, TDS and vanadium (a number of which only occur in the transition zone). Bedrock monitoring wells installed within the ash basin indicate only strontium detected greater than the background concentration. Boron is a key indicator of CCR groundwater impacts. Manganese and strontium also indicate impact at Mayo. Boron is detected at concentrations greater than the 2L beneath and downgradient of the ash basin. The area downgradient at which boron, 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 15-3 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx manganese, and strontium are detected at a concentration greater than PBTVs is interpreted as the leading edge of the CCR-derived plume moving downgradient from the source area. For the surficial flow zone, boron, manganese, and strontium were detected in monitoring wells screened in Crutchfield Branch alluvium downgradient of the ash basin and north of the Plant property boundary. The leading edge of the transition zone and bedrock plume is interpreted to be at/near the northern Plant property line. The surficial and transition zone flow units at Mayo, though impacted, are not vertically extensive. Impact to the bedrock flow unit is present in the upper 50 to 75 feet of fractured bedrock. Surface water data reflect that the majority of the flow in the former Crutchfield Branch stream valley is associated with engineered drainage from the ash basin. Boron concentrations are greatest proximate to the engineered toe drains and decrease toward the state line. Manganese concentrations are similarly consistent immediately downstream of the ash basin and begin to decline downstream; however, at the most downstream surface water location, manganese concentrations increase. The SCM will continue to be refined following evaluation of the completed groundwater models to be presented in the CAP and additional information obtained in subsequent data collection activities. 15.3 Interim Monitoring Program An Effectiveness Monitoring Program (EMP) is required by CAMA §130A-309.209 (b)(1)e. The EMP for the Mayo Plant is anticipated to begin once the basin closure and groundwater CAP have been implemented. In the interim, an IMP has been developed at the direction of NCDEQ. The CAP, and a proposed EMP, will be submitted at a future date. IMP Implementation 15.3.1 An IMP has been implemented in accordance with NCDEQ correspondence (NCDEQ, October 19, 2017; Appendix A) that provided an approved “Revised Interim Monitoring Plans.” Sampling will be conducted quarterly until approval of the CAP or as otherwise directed by NCDEQ. Groundwater samples will be collected using low-flow sampling techniques in accordance with the Low Flow Sampling Plan, Duke Energy Facilities, Ash Basin Groundwater Assessment Program, North Carolina, June 10, 2015 (Appendix G) conditionally approved by NCDEQ in a June 11, 2015 email with an attachment summarizing its approval conditions. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 15-4 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx Samples will be analyzed by a North Carolina certified laboratory for the parameters listed in Table 15-1. The table includes targeted minimum detection limits for each listed constituent. Analytical parameters and detection limits for each medium were selected so the results could be used to evaluate the effectiveness of a future remedy, conditions within the aquifer that may influence the effectiveness of the remedy, and migration of constituents related to the ash basin. Laboratory detection limits for each constituent are targeted to be at or below applicable regulatory values (i.e., 2L, IMAC, or 2B). Monitoring wells and surface water locations that will be sampled and monitored as part of the IMP, as approved in NCDEQ correspondence (NCDEQ, October 19, 2017; Appendix A), are included in Table 15-2. IMP Reporting 15.3.2 Currently, data summary reports comprised of analytical results received during the previous month are submitted to NCDEQ on a monthly basis. In addition, NCDEQ directed that an annual IMP report be submitted by April 30 of the following year of data collection. The reports shall include materials that provide “an integrated, comprehensive interpretation of site conditions and plume status.” The initial report was to be submitted to NCDEQ no later than April 30, 2018; however, the October 19, 2017 correspondence provides that the required date for an annual monitoring report will be extended to a date in 2018 to be determined later. 15.4 Preliminary Evaluation of Corrective Action Alternatives Closure of the ash basin is required by 2024 under CAMA (Intermediate Risk). The updated risk assessment (Section 12.0) has determined there is no imminent risk to human health or the environment due to groundwater, surface water, or sediment impacts. In three ash basin locations where soil samples could be collected, analytical results indicate a few, limited detections of COIs above PBTVs. If needed, groundwater and surface water can be remediated over time using a variety of approaches and technologies. Groundwater modeling has indicated closure by excavation compared to a cap-in-place closure does not substantively accelerate groundwater clean-up. For basin closure, bulk dewatering and reduction of infiltrating water will have the greatest positive impact on groundwater and surface water quality downgradient of the ash basin. Closure design can augment an overall groundwater corrective action scenario including cap-in-place or active groundwater remediation which will be evaluated in the CAP. Therefore, a “low” groundwater risk classification is recommended. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 15-5 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx This preliminary evaluation of corrective action alternatives is included to provide insight into the updated CAP preparation process. The preliminary evaluation is based on data available and the current understanding of regulatory requirements for the Site. It is assumed a source control measure of either implementation of an engineering capping system (cap-in-place) to minimize infiltration, or excavation of the ash within the ash basin, or a combination of the two, will be designed following completion of the risk classification process. Groundwater currently presents minimal , if any, risk to receptors. A “Low” risk classification and closure via a cap-in-place scenario are considered viable. Potential groundwater remedial strategies are being considered as part of the closure design. CAP Preparation Process 15.4.1 The CAP preparation process is designed to identify, describe, evaluate, and select remediation alternatives with the objective of bringing groundwater quality to levels that meet applicable standards, to the extent that the objective is economically and technologically feasible, in accordance with 2L .0106 Corrective Action. Sections (h), (i), and (j) regarding CAP preparation read as follows: (h) Corrective action plans for restoration of groundwater quality, submitted pursuant to Paragraphs (c), (d), and (e) of this Rule shall include: (1) A description of the proposed corrective action and reasons for its selection; (2) Specific plans, including engineering details where applicable, for restoring groundwater quality; (3) A schedule for the implementation and operation of the proposed plan; and (4) A monitoring plan for evaluating the effectiveness of the proposed corrective action and the movement of the contaminant plume. (i) In the evaluation of corrective action plans, the Secretary shall consider the extent of any violations, the extent of any threat to human health or safety, the extent of damage or potential adverse impact to the environment, technology available to accomplish restoration, the potential for degradation of the contaminants in the environment, the time and costs estimated to achieve groundwater quality restoration, and the public and economic benefits to be derived from groundwater quality restoration. (j) A corrective action plan prepared pursuant to Paragraphs (c), (d), or (e) of this Rule shall be implemented using a remedial technology demonstrated to provide the most effective means, taking into consideration geological and hydrogeological conditions at the contaminated site, for restoration of groundwater quality to the 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 15-6 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx level of the standards. Corrective action plans prepared pursuant to Paragraphs (c) or (e) of this Rule may request an exception as provided in Paragraphs (k), (l), (m), (r), and (s) of this Rule. To meet these requirements and to provide a comprehensive evaluation, it is anticipated that the CAP will include: Corrective action objectives and evaluation criteria Technology assessment Formulation of remedial action alternatives Analysis, modeling, selection, and description of selected remedial action alternative(s) Conceptual design elements, including identification of pre-design testing such as pilot studies Monitoring requirements and performance metrics Implementation schedule The following Site conditions significantly limit the effectiveness of a number of possible technologies. The COIs in groundwater flow primarily through the transition zone and upper fractured bedrock. The formations are very heterogeneous with anisotropic flow conditions. The preliminary screening of potential groundwater corrective action included: Source control by capping in place or excavation, and monitored natural attenuation, will be vital components to the CAP. Groundwater migration barriers. The lateral extent potentially required, along with the depth and heterogenitity of the transition zone and bedrock, may limit the feasibility of this technology. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 15-7 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx In situ chemical immobilization. This technology has not been demonstrated to be effective for the primary COI, boron. It may be applicable for other COIs. Permeable reactive barrier. Similar to in situ chemical immobilization, permeable reactive barrier technology has not been demonstrated to be effective for boron. Groundwater extraction. Groundwater extraction appears to be most likely and could potentially be a viable choice as a key element of groundwater corrective action in combination with source control and MNA. However, further analysis is required and will be addressed in the updated CAP. Potentially viable options will be further evaluated in the CAP with updated fate and transport and geochemical modeling. Summary 15.4.2 This preliminary evaluation of corrective action alternatives is intended to provide insight into the revised CAP preparation process, as outlined in 2L. It is based on data available and the current regulatory requirements for the Site. It addresses potentially applicable technologies and remedial alternatives. Potential approaches are based on the currently available information about Site hydrogeology and COIs. In general, three hydrogeologic units or zones of groundwater flow can be described for the Site: shallow/surficial zone, transition zone, and bedrock flow zone. The Site COIs include a list of common coal ash related constituents such as boron and manganese. If required, the potentially applicable technologies to supplement source control and MNA include groundwater extraction technologies such as conventional vertical wells, angle-drilled and horizontal wells. All of these extraction technologies could be augmented with fracturing of the bedrock formation. Migration barriers, in situ chemical immobilization, and permeable reactive barriers are also identified as potentially applicable remedial action alternatives. In the event that extracted groundwater may require treatment prior to discharge, several water treatment technologies for the relevant COIs would be evaluated, including pH adjustment, metals precipitation, ion exchange, permeable membranes, and adsorption technologies. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 15-8 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx The CAP will further evaluate basin closure options to reduce the potential impacts to human health or the environment; short- and long-term effectiveness, implementability, and potential for attenuation of contaminants; time and cost to achieve restoration; public and economic benefits; and compliance with applicable laws and regulations. The CAP evaluation process will be used to determine which approach, or combination of approaches, will be most effective. Modeling will also be used to evaluate the various options prior to selection. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 16-1 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx 16.0 REFERENCES ASTM. (2014). E1689-95: Standard Guide for Developing Conceptual Site Models for Contaminated Sites. ATSDR. (2012). Toxicological profile for Manganese. Atlanta: U.S. Department of Health and Human Services, Public Health Service. Ayotte, J. D., Gronbert, J. M., & Apodaca, L. E. (2011). Trace Elements and Radon in Groundwater Across the United States 1992-2003. U.S. Geological Survey Scientific Investigations Report 2011-5059. Bowman, J. D. (2010). The Aaron Formation: Evidence for a New Lithotectonic Unit in Carolina. North Central North Carolina. North Carolina State University. Raleigh, North Carolina: unpublished M.S. thesis. Bowman, J., Hibbard, J., & Miller, B. (2013, November 8). The Virgilina sequence redefined, north-central North Carolina, in One Arc, Two Arcs, Old Arc, New Arc – The Carolina Terrane in Central North Carolina. Carolina Geological Society. Butler, J., & Secor, D. (1991). The Central Piedmont, in the Geology of the Carolinas. In J. W. Horton, & V. A. Zullo (Eds.), The geology of the Carolinas: Carolina Geological Society fiftieth anniversary volume (1 ed.). Knoxville, TN: Univ. of Tennessee Press. Carpenter, A. P. (1976). Metallic mineral deposits of the Carolina Slate Belt. North Carolina. (84). Chapman, M. J., Cravotta, III, C. A., Szabo, Z., & Lindsey, B. D. (2013). Naturally occurring contaminants in the Piedmont and Blue Ridge crystalline-rock aquifers and Piedmont Early Mesozoic basin siliciclastic-rock aquifers, eastern United States, 1994– 2008. United States Geological Survey, Water Resources Investigations Report 00- 4286. Cox, J., Lundquist, G., Przyjazny, A., & Schmulbach, C. (1978). Leaching of boron from coal ash. Environmental Science & Technology, 12(6), 722-723. Cunningham, W. L., & Daniel, C. L. (2001). Investigation of ground-water availability and quality in Orange County, North Carolina. North Carolina: U.S. Dept. of the Interior, U.S. Geological Survey ; Branch of Information Services. Daniel, C. C., & Dahlen, P. R. (2002). Preliminary hydrogeologic assessment and study plan for a regional ground-water resource investigation of the Blue Ridge and Piedmont 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 16-2 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx provinces of North Carolina. Raleigh, North Carolina: U.S. GEOLOGICAL SURVEY Water-Resources Investigations Report 02–4105. Dennis, A. J., & Shervais, J. W. (1991). Arc rifting of the Carolina terrane in northwestern South Carolina. (Vol. 45). Geological Society of America. Dennis, A. J., & Shervais, J. W. (1996). The Carolina terrane in northwestern South Carolina: Insights into the development of an evolving island arc. Avalonian And Related Peri-Gondwana Terranes Of The Circum-North Atlantic, Special Paper 304, 237-156. (D. Nance, & M. Thompson, Eds.) Boulder, Colorado: Geological Society of America. Dewberry. (2016). Mayo Steam Station Phase II Potable Water Programmatic Evaluation - October 27, 2016. Roxboro, NC. Dicken, C., Nicholson, S., Horton, J., Foose, M., & Mueller, J. (2007). Preliminary integrated geologic map databases for the United States – Alabama, Florida, Georgia, Mississippi, North Carolina, And South Carolina. Version 1.1. Dudas, M. (1981). Long-term leachability of selected elements from fly ash. Environmental Science & Technology, 15(7), 840-843. Duke Energy. (2017). Retrieved October 20, 2017, from Duke Energy: https://www.duke-energy.com/_/media/pdfs/our-company/ash- management/duke-energy-ash-metrics.pdf?la=en Dzombak, D., & Morel, M. (1990). Surface complexation modeling: Hydrous ferric oxide. New York, NY: Wiley-Interscience Publication. EPRI. (1993). Physical and Hydraulic Properties of Fly Ash and Other By-Products From Coal Combustion. Palo Alto, CA. TR-101999: Electric Power Research Institute. EPRI. (1994). A Field and Laboratory Study of Solute Release from Sluices Fly Ash. Palo Alta, CA: Electric Power Research Institute. EPRI. (1995). Coal ash disposal manual: Third edition. Palo Alto, CA: Electric Power Research Institute, TR-104137. EPRI. (2005). Chemical Constituents in Coal Combustion Product Leachate: Boron. Palo Alto: Electric Power Research Institute. EPRI. (2006). Groundwater Remediation of Inorganic Constituents at Coal Combustion Product Management Site: Overview of Technologies, Focusing on Permeable Reactive Barriers. Palo Alto, CA: Electric Power Research Institute. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 16-3 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx EPRI. (2008a). Toxics Release Inventory. Chemical Profile: Arsenic. Palo Alto, CA: Electric Power Research Institute. EPRI. (2008b). Toxics Release Inventory. Chemical Profile: Barium. Palo Alto, CA: Electric Power Research Institute. EPRI. (2008c). Chemical Profile: Chromium. Electric Power Research Institute, Palo Alto, CA, and Hydro One Networks, Inc., Toronto, Canada: 2008. 1014622. EPRI. (2010). Comparison of coal combustion products to other common materials. Palo Alto, CA: Electric Power Research Institute, TR-1020556. Finkelman, R. (1995). Modes of occurrence of environmentally-sensitive trace elements in coal. In D. Swaine, & F. Goodzarzi, Environmental Aspects of Trace Elements in Coal (pp. 24-50). Kluwer Academic Publishers. Fleet, M. (1965). Preliminary investigations into the sorption of boron by clay minerals. Clay Minerals, 6(1): 3-16. Ford, R. G., W., R. T., & Puls, R. W. (2007). Monitored Natural Attenuation of Inorganic Contaminants in Ground Water. Cincinnati, OH: National Risk Management Research Laboratory, U.S. EPA. Freeze, R. A., & Cherry, J. A. (1979). Groundwater. Englewood Cliffs, NJ: Prentice-Hall. Freeze, R., & Witherspoon, P. (1967). Theoretical analysis of regional groundwater flow. 2. Effect of water-table configuration and subsurface permeability variation. Water Resources Research, 3(2): 623-624. Fruchter, J. S., Rai, D., & Zachara, J. M. (1990). Identification of solubility-controlling solid phases in a large fly ash field lysimeter. Environmental Science and Technology, 24(8), 1173-1179. Gillespie, E. (2013). Characterizing the Sources and Variability of Manganese in Well Water of the North Carolina Piedmont. International Conference on the Biogeochemistry of Trace Elements. Glover, L., & Sinha, A. (1973). The Virgilina deformation, a late precambrian to early cambrian (?) orogenic event in the central piedmont of Virginia and North Carolina. American Journal of Science, 273(A): 234-251. Goldberg, S. (1997). Reactions of boron with soils. Plant and Soil, 193: 35-48. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 16-4 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx Goldberg, S., Forster, H., Lesch, S., & Heick, E. (1996). Influence of anion competition on boron adsorption by clays and soils. Soil Science, 161 (2): 99-103. Goodarzi, F., Huggins, F., & Sanei, H. (2008). Assessment of elements, speciation of As, Cr, Ni and emitted Hg for a Canadian power plant burning bituminous coal. International Journal of Coal Geology, 74(1): 1-12. Haley & Aldrich. (2015). Evaluation of NC DEQ private well data, volumes 1 and 2. Harned, D., & Daniel, C. (1992). The transition zone between bedrock and regolith: Conduit for contamination? In Daniel, C.C., White, R., and Stone, P., eds., Groundwater in the Piedmont, Proceedings of a Conference on Ground Water in the Piedmont of the Eastern United States, Charlotte, N.C., Oct. 16-18, 1989. Clemson, SC: Clemson University (336-348). Harris, C., & Glover, L. (1985). The virgilina deformation: Implications of stratigraphic correlation in the Caroina Slate Belt. Carolina Geological Society, Guidebook for 1985 Annual Meeting. Harris, C., & Glover, L. (1988). The regional extent of the ca. 600 Ma Virgilina deformation: Implications for stratigraphic correlation in the Carolina terrane. The Geological Society of America, 100(2): 200-217. HDR and SynTerra. (2017). Statistical Methods for Developing Reference Background Concentrations for Groundwater and Soil at Coal Ash Facilities. HDR Engineering, Inc. and SynTerra Corporation. Heath, R. (1980). Basic elements of groundwater hydrology with reference to conditions in North Carolina. United States Geological Survey, Open-File Report: 80-44. Hem, J. D. (1985). Study and Interpretation of the Chemical Characteristics of Natural Water. United States Geological Survey, Water-Supply Paper 2254. Hibbard, J., Shell, G., Bradley, P., & Wortman, G. (1998). The Hyco shear zone in North Carolina and southern Virginia; implications for the Piedmont Zone-Carolina Zone boundary in the Southern Appalachians. American Journal of Science, 298(2): 85-107. Hibbard, J., Stoddard, E., Secor, D., & Dennis, A. (2002). The Carolina zone: Overview of Neoproterozoic to early Paleozoic Peri-Gondwanan terranes along the eastern flank of the southern Appalachians. Earth Science Reviews, 57(3): 299-339. Hurlbut, C. S. (1971). Dana's manual of mineralogy (18 ed.). John Wiley & Sons Inc. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 16-5 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx Izquierdo, M., & Querol, X. (2012). Leaching behaviour of elements from coal combustion fly ash : An overview. International Journal of Coal Geology, 94. 54-56. Johnson, A. (1967). Specific yield : compilation of specific yields for various materials (Vols. 1662-D). Washington: Geological Survey water-supply paper. Jones, K. B., & Ruppert, L. F. (2017, February). Leaching of trace elements from Pittsburgh coal mill rejects compared with coal combustion products from a coal- fired power plant in Ohio, USA. United States Geological Survey Bulletin, 171, 130- 141. Jurgens, B., McMahon, P., Chapelle, F., & Eberts, S. (2009). An Excel® workbook for identifying Redox processes in ground water. Retrieved from U.S. Geologic Survey Open-File Report. 2009-1004: Available at: https://pubs.usgs.gov/of/2009/1004/ Karamalidis, A., & Dzombak, D. (2010). Surface Complexation Modeling: Gibbsite. Hobboken, NJ: John WIley and Sons, Inc. Kunesh, C. (1978). Barium. (M. Grayson, & D. Eckroth, Eds.) Kirk-Othmer encyclopedia of chemical technology, Vol. 3, pp. 457–463. Laney, F. (1917). The geology and ore deposits of the Virgilina district of Virginia and North Carolina. Virginia Geological Survey and the North Carolina Geological and Economic Survey, Bulletin, (XIV). LeGrand, H. (1988). Region 21, Piedmont and Blue Ridge. In: J. Black, J. Rosenshein, P. Seaber, ed. Geological Society of America, 0-2, (pp. 201-207). LeGrand, H. (1989). A conceptual model of ground water settings in the Piedmont region, in groundwater in the Piedmont. In: Daniel C., White, R., Stone, P., ed. Ground Water in the Piedmont of the Eastern United States (pp. 317-327). Clemson, SC: Clemson University. LeGrand, H. (2004). A master conceptual model for hydrogeological site characterization in the Piedmont and Mountain Region of North Carolina: A guidance manual. North Carolina Department of Environment and Natural Resources, Division of Water Quality, Groundwater Section, Raleigh, NC, 55. Martell, A. E., & Smith, R. M. (2001). Critical Stability Constants. National Institute of Standards. Miao, Z., Brusseau, M. L., Carroll, K. C., & others. (2012). Environ Geochem Health. 34:539. https://doi.org/10.1007/s10653-011-9423-1. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 16-6 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx Miner, S. (1969). Air pollution aspects of barium and its compounds. Bethesda: Litton Systems. NCDENR. (2015). Classifications and water quality standards applicable to the surface waters of North Carolina (pending approval of 2007-2014 Triennial review). Raleigh: North Carolina Administrative Code Title 15A, Subchapter 02B. NCDHHS. (2010a). Concentration of Barium Detected in NC Private Well Water, Average 1998-2010 and Average 2010. Well Water & Health. University of North Carolina Superfund Research Program. NCDHHS. (2010b). Concentration of Iron Detected in NC Private Well Water, Average 1998-2010 and Average 2010. Well Water & Health. University of North Carolina Superfund Research Program. NCDNRCD. (1985). Geological map of North Carolina. North Carolina Geological Survey, Division of Land Resources. Niswonger, R., Panday, S., & Motomu, I. (2011). MODFLOW-NWT, A Newton formulation for MODFLOW-2005. In Techniques and Methods 6-A37 (pp. 1-44). Reston, Virginia: Chapter 37 of Section A, Book 6: U.S. Geological Survey. NOAA. (2013). Regional climate trends and scenarios for the U.S. national climate assessment. Part 2. Climate of the southeast U.S. Washington, D.C.: National Oceanic and Atmospheric Administration - U.S. Department of Commerce. Parker, R. (1967). Chapter D. composition of the Earth's crust. Geological survey paper 440-D. In Data of Geochemistry. 6th ed. Washington, D.C.: U.S. Government Printing Office, 1967. Polizzotto, M., Amoozegar, A., Austin, R., Bolich, R., Bradley, P., Duckworth, O., et al. (2015). Surface and Subsurface Properties Regulating Manganese Contamination of Groundwater in the North Carolina Piedmont. Water Resources Research Institute of The University of North Carolina. Pollock, J., Hibbard, J., & Sylvester, P. (2010). Depositional and tectonic setting of the Neoproterozoic-early Paleozoic rocks of the Virgilina sequence and Albemarle Group, North Carolilna. In R. Tollo, M. Batholomew, J. Hibbard, & P. Karabinos, From Rodinia to Pangea: The Lithotectonic Record of the Appalachian Region (pp. 1-34). Geological Society of America, 206. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 16-7 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx Robson, S. (1993). Techniques for estimating specific yield and specific retention from grain- size data and geophysical logs from clastic bedrock aquifers. U.S. Geological Survey, Water-Resources Investigation Report 93-4198, 19p. Rogers, J. (2006). Chapter 2 - The Carolina Slate Belt. In V. Steponaitis, J. Irwin, T. McReynolds, & C. Moore, Stone Quarries and Sourcing in the Carolina Slate Belt (pp. 10-15). Chapel Hill, NC: University of North Carolina Research Laboratories of Archaeology, Research Report 25. Ruhl, L., Vengosh, A., Dwyer, G. S., Hsu-Kim, H., Schwartz, G., Romanski, A., et al. (2012, September 30). The Impact of Coal Combustion Residue Effluent on Water Resources: A North Carolina Example. Environmental Science and Technology, 12226-12233. Samson, S., Hibbard, J., & Wortman, G. (1995). Nd isotopic evidence for juvenile crust in the Carolina terrance, southern Appalachians. Contributions to Mineralogy and Petrology; 121(2), 171-184. Schmidt, H. (1962). Equilibrium diagrams for minerals at low temperature and pressure. The Geological Club of Harvard, 199. Smith, L. A., Means, J. L., Chen, A., & others. (1995). Remedial Options for Metals- Contaminated Sites. Boca Raton, FL: Lewis Publishers. Smith, S. M. (2016). National Geochemical Database. Retrieved October 20, 2017, from USGS: https://pubs.usgs.gov/of/1997/ofr-97-0492/ State Climate Office. (2017). Retrieved October 2017, from www.nc- climate.ncsu.edu/climate/ncclimate.html SynTerra. (2014a). Drinking water well and receptor survey - Mayo Steam Electric Plant - September 2014. Roxboro, NC. SynTerra. (2014b). Supplement to drinking water well and receptor survey - Mayo Steam Electric Plant - November 2014. Roxboro, NC. SynTerra. (2014c). Proposed groundwater assessment work plan for Mayo Steam Electric Plant (Revision 1) - December 30, 2014. Roxboro, NC. SynTerra. (2015a). Comprehensive site assessment report - Mayo Steam Electric Plant - September 2, 2015. Roxboro, NC. SynTerra. (2015b). Corrective action plan - part 1: Mayo Steam Electric Plant - December 1, 2015. Roxboro, NC. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 16-8 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx SynTerra. (2016a). Corrective action plan - part 2: Mayo Steam Electric Plant - February 29, 2016. Roxboro, NC. SynTerra. (2016b). Comprehensive site assessment, supplement 1 - Mayo Steam Electric Plant - July 7, 2016. Roxboro, NC. SynTerra. (2016c). Update to drinking water well and receptor survey - Mayo Steam Electric Plant - September 2016. Roxboro, NC. Urey, H., & Mem, R. (1953). On the concentration of certain elements at the earth's surface. In Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences (pp. 281-292). The Royal Society Publishing, 219(1138). USEPA. (1997). Ecological risk assessment guidance for superfund: Process for designing and conducting ecological risk assessments. Edison, NJ: U.S. Environmental Protection Agency, Environmental Response Team. EPA 540-R-97-006. USEPA. (2003). Drinking Water Advisory: Consumer Acceptability Advice and Health Effects Analysis on Sulfate. Washington: U.S. Environmental Protection Agency, EPA 822- R-03-007. USEPA. (2012). Sulfate in Drinking Water. Retrieved from EPA: http://water.epa.gov/drink/contaminants/unregulated/sulfate.cfm USEPA. (2015). Disposal of Coal Combustion Residual From Electric Utilities: Final Rule - April 17, 2015. Code of Federal Register, Vol. 80(No. 74). USEPA. (2017a). National recommended water quality criteria for aquatic life. Retrieved October 20, 2017, from EPA: https://www.epa.gov/wqc/national-recommended- water-quality-criteria-aquatic-life-criteria-table#altable USEPA. (2017b). Secondary Drinking Water Regulations: Guidance for Nuisance Chemicals. Retrieved October 2017, from EPA: https://www.epa.gov/dwstandardsregulations/secondary-drinking-water- standards-guidance-nuisance-chemicals USEPA. (2017c). Watersense. Retrieved October 2017, from EPA: http://www.epa.gov/watersense/pubs/indoor.html USEPA. (October 2007). Monitored natural attenuation of inorganic contaminants in ground water technical basis for assessment, volume I. United States Environmental Protection Agency. EPA/600/R-07/139. USGS. (1973). United States Mineral Resources. United States Government Printing Office. 2017 Comprehensive Site Assessment Update October 2017 Mayo Steam Electric Plant SynTerra Page 16-9 P:\Duke Energy Progress.1026\105. Mayo Ash Basin GW Assessment Plan\CSA_UPDATE_2017\CSA Sup 2\Text\FINAL Mayo CSA October 2017.docx WHO. (1996). Guidelines for drinking-water quality: Health criteria and other supporting information. (2nd, Ed.) Vol. 2. Wilkins, J., Shell, G., & Hibbard, J. (1995). Geologic contrasts across the central Piedmont suture in north-central North Carolina. Field Trip Guide for the 1995 Carolina Geological Society Annual Meeting: Geology of the Western Part of the Carolina Terrance in Northwestern South Carolina, 37, pp. 25-32. Wortman, G., Samson, S., & Hibbard, J. (2000). Precise U-Pb Zircon constraints on the earliest magmatic history of the Carolina Terrance. The Journal of Geology, 108(3), 321-338. Zhang, D. (2013). Ultra-supercritical coal power plants: Materials, technologies and optimisation. Sawston, Cambridge, UK: Woodhead Publishing.