Loading...
HomeMy WebLinkAboutMarshall Attachments_20160418Attachments Expert Report of Douglas Cosler, Ph.D., P.E. Expert Report of Douglas J. Cosler, Ph.D., P.E. Chemical Hydrogeologist Adaptive Groundwater Solutions LLC Charlotte, North Carolina Marshall Steam Station Ash Basin Terrell, North Carolina April 18, 2016 Introduction Site Backqround The Marshall Steam Station is a four -unit, coal-fired generating station owned by Duke Energy and located on a 1,446-acre site on the west bank of Lake Norman near the town of Terrell in Catawba County, North Carolina. Coal combustion residuals ("coal ash") have historically been disposed in a single -cell ash basin impoundment located north of the power plant. The ash basin was constructed in 1965 by building an earthen dike at the confluence of Holdsclaw Creek and the Catawba River (now Lake Norman) and is generally located in historical depressions formed from Holdsclaw Creek and small tributaries that discharged into the creek. Two unlined ash landfill units (Marshall dry ash landfill) are located adjacent to the east (Phase 1) and northeast (Phase II) portions of the ash basin. A flue gas desulfurization (FGD) landfill containing bottom ash and various other types of waste materials is located to the west of the ash basin and is constructed with an engineered liner system. Industrial Landfill No. 1 (containing fly and bottom ash and various other types of waste materials) is located adjacent to the northern portion of the ash basin and contains a liner and leachate collection system. The demolition landfill (construction and demolition waste) is also located adjacent to the northern portion of the ash basin (directly north of the dry ash landfill, Phase II). The photovoltaic farm structural fill (PV structural fill) is constructed of fly ash and located adjacent to and partially on top of the northwest portion of the ash basin. Duke Energy performed voluntary groundwater monitoring at the site from November 2007 to October 2011 (nine sampling events) and NPDES permit -required compliance monitoring starting in February 2011 (sampling three times per year). Recent groundwater sampling results at Marshall indicate exceedances of 15A NCAC 02L.0202 Groundwater Quality Standards (2L Standards). In response to this, the North Carolina Department of Environmental Quality (NC DEQ) required Duke Energy to perform a groundwater assessment at the site and prepare a Comprehensive Site Assessment (CSA) report. The Coal Ash Management Act of 2014 (CAMA) also required owners of surface impoundments containing coal combustion residuals (CCR) to conduct groundwater monitoring and assessment and prepare a CSA report. The September 2015 CSA prepared by HDR Engineering, Inc. of the Carolinas (HDR) for the Marshall site determined that ash handling and storage at the Marshall site have impacted soil and groundwater beneath and downgradient from the ash basin. The CSA report identified Constituents of Interest (COI) considered to be associated with potential impacts to soil and groundwater from the ash basin and assessed COI concentration distributions in soil, groundwater, and seeps. CAMA also requires the submittal of a Corrective Action Plan (CAP); the CAP for the Marshall site consists of two parts. CAP Part 1 (submitted to DEQ in December 2015) provides a summary of CSA findings, further evaluation and selection of COI, a site conceptual model (SCM), the development of 2 groundwater flow and chemical transport models of the site, presentation and analysis of the results of the modeling, and a quantitative analysis of groundwater and surface water interactions. The CAP Part 2 contains proposed remedial methods for achieving groundwater quality restoration, conceptual plans for recommended corrective action, proposed future monitoring plans, and a risk assessment. Information Reviewed My opinions are based upon an analysis and technical review of (i) hydrogeologic and chemical data collected at the Marshall site; (ii) the analyses, interpretations, and conclusions presented in site -related technical documents and reports; (iii) the groundwater flow and chemical transport models constructed for the site (including model development, calibration, and simulations of remedial alternatives); (iv) the effectiveness of proposed remedial alternatives to achieve groundwater quality restoration; and (v) proposed future site monitoring. These opinions are subject to change as new information becomes available. As a basis for forming my opinions I reviewed the following documents and associated appendices: (1) Comprehensive Site Assessment Report, Marshall Steam Station Ash Basin (September 8, 2015); (2) Corrective Action Plan, Part 1, Marshall Steam Station Ash Basin (December 7, 2015); (3) Corrective Action Plan, Part 2, Marshall Steam Station Ash Basin (March 3, 2016); (4) Miscellaneous historical groundwater and soil concentration data for the Marshall site collected prior to the CSA; and (5) Specific references cited in and listed at the end of this report. Professional Qualifications I have advanced graduate degrees in Hydrogeology (Ph.D. Degree from The Ohio State University) and Civil and Environmental Engineering (Civil Engineer Degree from the Massachusetts Institute of Technology), and M.S. and B.S. degrees from Ohio State in Civil and Environmental Engineering. I have 36 years of experience as a chemical hydrogeologist and environmental engineer investigating and performing data analyses and computer modeling for a wide variety of projects. These projects include: investigation, remediation, and regulation of Superfund, RCRA, and other hazardous waste sites involving overburden and bedrock aquifers; ground water flow and chemical transport model development; natural attenuation/biodegradation assessments for chlorinated solvent and petroleum contamination sites; volatile organic compound vapor migration and exposure assessment; exposure modeling for health risk assessments; hydrologic impact assessment for minerals and coal mining; leachate collection system modeling and design for mine tailings disposal impoundments; and expert witness testimony and litigation support. I also develop commercial groundwater flow and chemical transport modeling software for the environmental industry. 3 The types of sites I have investigated include: landfills, mining operations, manufactured gas plants, wood -treating facilities, chemical plants, water supply well fields, gasoline and fuel oil storage/delivery facilities, nuclear waste disposal sites, hazardous waste incinerators, and various industrial facilities. I have investigated the following dissolved, nonaqueous-phase (LNAPL/DNAPL), and vapor -phase contaminants: chlorinated solvents, various metals, gasoline and fuel oil constituents, wood -treating products, coal tars, polychlorinated biphenyls, pesticides, dioxins and furans, phenolic compounds, flame retardants (PBDE), phthalates, radionuclides, and biological constituents. Summary of Opinions The following is a brief summary of the opinions developed in my report: • A total of 21 Compliance Boundary groundwater samples exceeded North Carolina groundwater standards for these COI: antimony, boron, chromium, cobalt, iron, manganese, total dissolved solids, and vanadium. Of these 21 exceedances, 19 were greater than the proposed provisional background concentrations by HDR, which exaggerate background levels; • The statistical analyses of historical, shallow -aquifer background groundwater concentrations at the Marshall site (monitoring well MW-4) are invalid due to the characteristically slow rate of COI migration in groundwater; • There is a significant risk of chemical migration from the ash basin to neighboring private and public water supply wells in fractured bedrock; • Major limitations of the CAP Parts 1 and 2 groundwater flow and chemical transport models prevent simulation and analysis of off -site migration; • The CAP Closure Scenario simulations greatly underestimate (by factors of 10 or more) the time frames required to achieve meaningful groundwater concentration reductions in response to remedial actions; • For either the Existing Condition or Cap -in -Place Scenario groundwater concentrations of many coal -ash constituents much higher than background levels will continue to exceed North Carolina groundwater standards at the Compliance Boundary because saturated coal -ash material and secondary sources will remain in place; • Source -area mass removal included in the Excavation Scenario results in COI groundwater concentration reductions that are at least a factor of 10 greater compared to Cap -in -Place at many locations. Additional excavation of secondary sources would further accelerate concentration reductions; • CSA data show 22 exceedances of groundwater standards in bedrock inside the Compliance Boundary. However, the CAP Closure Scenarios do not address off -site chemical migration control in the fractured bedrock aquifer; 4 • Due to an incorrect boundary -condition representation of the active ash basin, the CAP models underestimate by a factor of three or more both the mass loading of COI into Lake Norman and the corresponding Lake Norman water concentrations (attributable to coal ash ponds) estimated by the groundwater/surface-water mixing model; • The CAP Part 2 geochemical modeling and monitored natural attenuation (MNA) evaluations do not provide the required quantitative analyses of COI attenuation rates necessary to support MNA as a viable corrective action. The CAP 2 chemical transport modeling, which included attenuation by sorption, demonstrated that MNA is not an effective remedial option for several COI (e.g., antimony, beryllium, boron, chromium, cobalt, hexavalent chromium, thallium, and vanadium); and • Future Compliance Monitoring at the Marshall site should include much more closely -spaced Compliance Wells to provide more accurate detection, and groundwater sampling frequency should be re-evaluated to allow valid statistical analyses of concentration variations. Hydrogeology of the Marshall Site Introduction The groundwater system at the Marshall site is an unconfined, connected system consisting of three basic flow layers: shallow, deep, and fractured bedrock. The shallow and deep layers consist of residual soil, saprolite (clay and coarser granular material formed by chemical weathering of bedrock), and weathered fractured rock (regolith). A transition zone at the base of the regolith is also present and consists of partially-weathered/fractured bedrock and lesser amounts of saprolite. The single -cell ash basin impoundment overlies native soil and was constructed in 1965 by building an earthen dike at the confluence of Holdsclaw Creek and the Catawba River (now Lake Norman). Most of the coal ash lies below the groundwater table and is saturated. Groundwater flow through saturated coal ash and downward infiltration of rainwater through unsaturated coal ash leach COI into the subsurface beneath the basin and via seeps through the embankments. As described by HDR, groundwater flow in all three layers within the site boundary is generally from the northwest and north to the southeast toward Lake Norman. Vertical groundwater flow between the three layers also occurs, and surface water ponding in the ash basin effects flow directions locally. On the downgradient eastern site boundary the CSA and CAP Parts 1 and 2 investigations assumed that all groundwater (overburden and bedrock aquifers) discharges into Lake Norman. However, these studies did not collect hydrogeologic data or perform data analyses or groundwater flow modeling to support this assumption. The CSA and CAP Parts 1 and 2 also did not analyze potential changes to site groundwater flow directions, or the risk of off -site migration of COI in the overburden or bedrock aquifers, caused by 5 groundwater extraction from numerous private and public water supply wells located close to the site boundaries. My report begins with a discussion of significant errors in CSA data analysis and conceptual model development that contradict HDR's interpretation of three-dimensional groundwater flow patterns at the Marshall site. This is followed by a presentation and discussion of measured exceedances of North Carolina groundwater standards at multiple locations on the ash basin compliance boundary. I then address several limitations of the CAP Parts 1 and 2 groundwater flow and chemical transport models and identify various model input data errors. Finally, I present my evaluations of the CAP Closure Scenario simulations and provide my opinions regarding the effectiveness of various remedial alternatives for restoring groundwater quality to North Carolina standards. Errors in Hydraulic Conductivity Test Analyses Background In the CSA and CAP reports HDR provides interpretations and conclusions regarding the horizontal and vertical variations of groundwater flow directions and rates, and the fate and transport of COI dissolved in groundwater. The most important site -specific parameter that controls these time -dependent flow and transport mechanisms is the hydraulic conductivity (also referred to as "permeability") of the underlying soils and fractured bedrock (Bear, 1979). Hydraulic conductivity (length/time) is a media -specific measure of the rate at which water can flow through a porous (soil) or fractured (bedrock) porous medium. Groundwater flow and chemical transport rates are directly proportional to the product of hydraulic conductivity and the hydraulic gradient (hydraulic head difference between two points divided by the separation distance; e.g., the water table elevation slope). Therefore, accurate measurement of hydraulic conductivity is critical for understanding the current and future distributions of COI in soil and groundwater and for evaluating the effectiveness (e.g., cleanup times) of alternative remedial measures. In addition, the contrast in hydraulic conductivity between adjacent hydrogeologic units is the key factor in determining three-dimensional groundwater flow directions and the ultimate fate of dissolved COI. For example, at the Marshall site accurate measurement of hydraulic conductivity is critical in evaluating the potential for: downward chemical migration into the fractured bedrock unit, off -site COI migration in the overburden (soil) or fractured bedrock aquifers, and groundwater flow and COI transport into or beneath Lake Norman. A slug test is one of the standard field methods for measuring hydraulic conductivity (K) using a soil boring or installed monitoring well. Slug tests were performed in most of the overburden and bedrock wells at the Marshall site. In this test the static water level in the open hole (boring) or well casing is N. suddenly increased or decreased and the resulting transient change in water level is recorded. Two commonly -used techniques for quickly changing the water level are the introduction (increases the water level) and removal (decreases the water level) of a solid rod, or "slug" into the boring or well casing. These tests are called "falling -head" and "rising -head" tests, respectively. Higher rates of water -level recovery correspond to higher values of K. The measurements of water level versus time are analyzed using mathematical models of the groundwater flow hydraulics and information regarding the well installation (e.g., length of the slotted monitoring well screen and well casing diameter) to compute an estimate of K. As discussed below, HDR made significant errors in all of their analyses of field slug test data. Their analysis errors caused the reported (CSA report) slug test hydraulic conductivity values to be as large as a factor of 1.5 (almost 50 percent) smaller than the correct K values. I discuss the impacts of these analysis errors on HDR's groundwater flow and chemical transport assessments and the CAP modeling later in my report. Overburden Slug Tests HDR analyzed all of the CSA overburden slug tests in shallow and deep wells with the Bouwer-Rice (1976) method using vertical anisotropy, Av = Kh,,n onrat 1 Kverticat , values that are as large as a factor of 20 lower than the values presented in the CSA report (e.g., compare geometric mean values in CSA Tables 11-9 and 11-10) and used in the CAP modeling (e.g., CAP 1 report Appendix C, Table 2), where K is hydraulic conductivity. Comparing CSA Tables 11-9 and 11-10, the measured Av for overburden soil units ranges from 2 to 50. In the calibrated CAP flow model Av = 10 for overburden soil. However, the Bouwer-Rice slug test analyses assumed Av = 2 for every monitoring well (CSA Appendix H). Using the measured M1 hydrostratigraphic vertical anisotropy, Av = 35, increases all of the M1 CSA overburden hydraulic conductivity values (CSA Table 11-3) by almost 50 percent (factor of 1.5), depending on how the slug -test radius of influence was computed. If the CAP 1 flow model results (Av = 10) are used in the Bouwer-Rice analyses all of the measured overburden hydraulic conductivity values increase by almost 30 percent (factor of 1.3). Since most of the reported overburden K values in the CSA report are up to 50 percent too low, the actual average chemical transport rates in overburden soils are up to 50 percent greater than reported. This site -wide data reduction error also affects the CAP flow and transport model calibrations. For example, the transport model developers significantly reduced laboratory measurements of the soil -water partition coefficient, Kd, for various COI during the transport model calibration based on comparisons of observed and simulated chemical migration rates. However, if the correct (i.e., higher) overburden K values had been used in the model calibration the Kd values would not have been reduced as much (compared to laboratory values). The reason for this is, assuming linear equilibrium partitioning of COI with soil, the 7 chemical migration rate is proportional to K/Kd (except for Kd << 1). The CAP 1 and 2 transport model history matching indicated that the simulated transport rate was too low, so HDR reduced the model Kd. In other words, the reductions in calibrated Kd values would not have been as great if the correct (higher) K values were used in the first place. As discussed below, the CAP Part 1 transport modeling used Kd values that were typically factors of 10 - 100 smaller than the measured site -specific Kd's reported in CAP Appendix D. In contrast, the CAP Part 2 transport modeling used Kd values that are generally a factor of about 10 larger than the CAP 1 values; however, the CAP 2 Kd's remain on the order of 10 times smaller than the measured site -specific Kd's for many COI. COI sorption to soil is important because, as discussed later, aquifer cleanup times (i.e., chemical flushing rates) are generally proportional to the chemical retardation factor, which is directly proportional to Kd , except when Kd << 1 (Zheng et al., 1991). Groundwater Flow In the CSA and CAP reports HDR made several critical assumptions, not supported by data, regarding the horizontal and vertical groundwater flow directions near the boundaries of the Marshall site which impacted their conclusions regarding the ultimate discharge locations for site groundwater and dissolved COI. Two examples discussed in this section are (i) the relationship between site groundwater and Lake Norman and (ii) groundwater flow directions and the potential for offsite migration of COI. Lake Norman and the LeGrand Conceptual Model All shallow, deep, and bedrock groundwater in the downgradient areas of the site was apparently assumed to discharge into either Lake Norman or the unnamed stream located close to the northeastern side of the ash basin before site -specific hydrogeologic data were analyzed. The rationale that HDR used is a generalized, theoretical conceptual model (LeGrand, 2004) that relies on land -surface topographic elevations rather than actual groundwater flow data (i.e., hydraulic head measurements); references to this theory were made at numerous points in the CSA and CAP reports. However, HDR did not present any site -specific data analyses or groundwater flow modeling that would support this assumption in either report. In fact, as discussed below, the boundary conditions for the CAP Parts 1 and 2 flow models effectively forced downgradient site groundwater in the shallow, deep, and bedrock aquifers to discharge into Lake Norman and the unnamed stream. HDR continued to state this assumption in the CAP 2 report (e.g., Section 3.3.2) even though strong measured downward groundwater flow components from deep overburden to the fractured bedrock unit exist in the southeastern part of the site next to Lake Norman. In the CAP 2 report Executive Summary (page 2) HDR also claims that "Groundwater at the MSS site generally flows .... to the south and southeast beneath the source areas toward Lake Norman, which serves as the hydrologic boundary downgradient of the CCR source areas". I demonstrate below that this interpretation is incorrect because F:3 CSA and CAP 2 hydraulic head data show that groundwater and dissolved COI in the fractured bedrock aquifer flow off -site in the southeastern portion of the Marshall site toward neighboring water supply wells The LeGrand (2004) guidance document presents a general discussion of groundwater flow patterns that may occur near streams in the Piedmont and Mountain Region of North Carolina based on ground surface elevations (i.e., site topography and surface watershed boundaries). However, surface water and groundwater watersheds commonly do not coincide (Winter et al., 2003). Further, groundwater flow patterns and rates in bedrock have been found to be poorly related to topographic characteristics (Yin and Brook, 1992). LeGrand does not present or derive any mathematical equations or quantitative relationships for groundwater flow near rivers or streams and does not address areally-extensive waterbodies such as Lake Norman. The author emphasizes that site -specific data must be collected in order to correctly evaluate river inflow or outflow. In strong contrast to the LeGrand generalizations, numerous detailed and sophisticated mathematical (analytical and numerical) river -aquifer models and highly -monitored field studies have been published in the scientific and engineering literature in the past several decades. What these investigations and applied hydraulic models show is that the water flow rate into or out of a river, stream, or lake and the depth of hydraulic influence within an underlying aquifer are highly sensitive to several factors, including: the transient river/lake water surface elevation and slope; river/lake bed topography; bed permeability and thickness; horizontal and vertical permeability (and thickness) of the different hydrogeologic units underlying the river/lake; transient horizontal and vertical hydraulic head variations in groundwater beneath and near the river/lake; and groundwater extraction rates and screen elevations for neighboring pumping wells (e.g., Simon et al. 2015; McDonald and Harbaugh, 1988; Bear, 1979; Hantush, 1964). The CSA investigation did not: measure the bed permeability or thickness of Lake Norman or the unnamed stream; characterize the lake/river bathymetry; monitor transient water surface elevation variations at more than one location (one average value was used); collect lake/river bed hydraulic gradient data; measure horizontal or vertical overburden or bedrock permeability beneath Lake Norman; characterize the geology beneath Lake Norman; measure hydraulic heads in the overburden or bedrock beneath Lake Norman or the stream; or consider the hydraulic effects of groundwater extraction from nearby water supply wells [e.g., water supply wells NC0118676 (Duke Energy), NC0118736 (Old Country Church), and NC0118622 (Midway Restaurant and Marina), located close to the southern section of the ash basin compliance boundary; and the numerous private homes located south and west of the site; see CSA Figures 4-1 and 4-5]. Well construction and well yield (gallons per minute) data for these wells are listed in Table 1 of the 2014 receptor survey report (HDR, 2014). Much of the CSA data contradict the LeGrand hypothesis at the Marshall site. For example, downward flow components from deep to bedrock wells were measured at monitoring clusters AB-5 and AB-6 M located in the southeastern portion of the site close to Lake Norman during the CSA and CAP 2 (compare CSA Figures 6-6 and 6-7 and CAP 2 Figures 2-3 and 2-4). In addition, my corrected contouring of bedrock hydraulic head data shows that groundwater flow in the bedrock aquifer is off -site and to the south/southeast in this portion of the site (see next section). As discussed in other parts of my report, HDR assumed in the CSA and the CAP Parts 1 and 2 modeling that groundwater flow in all three hydrogeologic units is directly west to east near the southern part of the Compliance Boundary and used a no -flow boundary in this model area, which prevents off -site flow. The CSA and CAP 1 and 2 also did not evaluate the effect of the very large constant hydraulic head in the ash basin impoundment (30 feet greater than the Lake Norman water level and the static bedrock hydraulic head in this area) as it relates to vertical groundwater flow from overburden to bedrock and COI transport from ash basin water to underlying groundwater. Maps of vertical hydraulic gradient variations (e.g., contour maps) were not generated for the CSA or CAP 2, and HDR did not discuss the significance of downward hydraulic gradients from overburden to bedrock, or from the impounded ash basin water to the aquifer system. These downward groundwater flow measurements are consistent with the hydraulic conductivities of the bedrock and overburden being of similar magnitude (CSA Table 11-9). Groundwater Flow Directions The CSA assumptions and analysis errors discussed above have had a strong effect on: the Conceptual Model development; the site hydrogeologic and COI transport assessment; the construction/calibration of the CAP flow and transport models; and the simulations of CAP Close Scenarios. The hydrogeologic assumptions should have been carefully evaluated and tested during the performance of the CSA and as part of the CAP groundwater flow model construction and calibration to determine whether they were valid. Instead, the hypotheses appear to have effectively guided the model development and led to inaccurate interpretations. A good example of this is the bedrock aquifer hydraulic head and flow direction map developed by HDR (CSA Figure 6-7). The hydraulic head contours and inferred bedrock flow directions in Figure 6-7 are inaccurate because HDR did not account for the hydraulic impacts of three very important influences on horizontal and vertical groundwater flow in the fractured bedrock aquifer: the ash basin impoundment, Lake Norman, and groundwater extraction from off -site water supply wells (e.g., the wells mentioned above). First, the 780-foot head contour is incorrect because it cuts through the middle of the ash basin where the constant hydraulic head is approximately 790 feet. Drawn correctly, the 780-foot head contour should mirror the northeastern boundary of the ash basin as shown in the deep aquifer map (CSA Figure 6-6 and CAP flow model Figure 11 of Appendix C). Second, the CSA bedrock flow map does not utilize the hydraulic information (specifically, the groundwater/surface-water interaction) associated with the Lake Norman shoreline area to the east and southeast of the Marshall site (e.g., CSA Figure 4-1). For example, a U.S. Geological Survey study of overburden and bedrock groundwater flow near another part 10 of the Lake Norman shoreline (Pippin et al., 2008) determined that heads in bedrock were generally less than a few feet greater than the Lake Norman water surface elevation, depending on how much rainfall had occurred during the days and weeks preceding the measurement event (e.g., Figure 25 in the USGS report). At the Marshall site, water -level measurements in monitoring well GWA-1 BR are consistent with this pattern. Similarly, static water levels in the three bedrock pumping wells to the south (Duke Energy, Old Country Church, and Midway Restaurant and Marina) are strongly influenced by hydraulic interaction with Lake Norman. As shown in CSA Figure 4-1 these wells are located very close to the Lake Norman shoreline. HDR (2014; Table 1) reports that the well yield for the Duke Well is 30 gallons per minute (gpm), and Duke Energy has reported average annual water usage rates of about 10 gpm for this well in annual water use reports submitted to the North Carolina Division of Water Resources (e.g., McGary, 2004). The reported well yields for Old Country Church and Midway Restaurant and Marina are 40 gpm and 20 gpm, respectively. To better understand the effects of these extraction wells on groundwater flow in the bedrock aquifer I estimated hydraulic head reductions, relative to the Lake Norman static level, as a function of distance from a pumping well using the exact mathematical model of Hantush (1964, Equation 73). This analytical groundwater flow model predicts the steady-state, pumping -induced drawdown [i.e., reduction in hydraulic head from static (non -pumping) conditions] in a confined aquifer (bedrock) overlain by a "leaky layer", which I model as the overlying overburden aquifer (using the overburden aquifer vertical hydraulic conductivity, as represented by data in CSA Table 11-9) and the Lake Norman water surface elevation. I assumed a bedrock aquifer thickness of 100 feet, which is much larger than the underestimated value used in the CAP models (i.e., 50 feet) but also is consistent with regional studies of this bedrock aquifer system (Daniel et al., 1989). The cited depths of the Duke and Midway Marina supply wells (HDR, 2014) are about 400 feet and 160 feet below the bedrock surface, respectively (540 and 245 feet below ground surface). For the overlying confining unit I used representative CSA values: a 50-foot thick overburden aquifer with a 0.045 feet/day vertical permeability. The following bedrock drawdown calculations depend on the product of bedrock permeability and thickness (i.e., transmissivity); I varied permeability by a factor of seven, which has a similar hydraulic effect on drawdown as varying the thickness by this amount. Therefore, my calculations also represent sensitivity analyses of both the thickness and the permeability of the fractured bedrock unit. I used a 30-gpm pumping rate, which is representative of the well yields for any of the three supply wells. The Hantush model is conservative (i.e., underestimates drawdown) as applied to this analysis because the model assumes that Lake Norman covers the entire area; in actuality the lake overlies only parts of the northeastern and southeastern quadrants of the Duke Energy supply well zone of influence. Figure 1 is a graph of the computed steady-state drawdown versus distance for three values of bedrock hydraulic conductivity (Kb,): 0.68 (mean site value from CSA Table 11-9), 2.0, and 4.7 feet/day 11 (measured value for well GWA-1 BR, which is the Marshall -site bedrock well located closest to the Duke Energy pumping well; CSA Table 11-3). The estimated drawdown ranges from 14 feet (higher Kb,) to 85 feet at the pumping well, and the radius of influence (e.g., drawdown = 0.01 feet) generally varies from 1800 feet (lower Kb,) to approximately 3,400 feet. Note that the distance from the Duke Energy pumping well to monitoring well GWA-1 BR is about 2,000 feet. Therefore, these drawdown analyses indicate that groundwater extraction from any of the supply wells NC0118676 (Duke Energy), NC0118736 (Old Country Church), or NC0118622 (Midway Restaurant and Marina) will likely influence bedrock groundwater flow directions in the southeastern portion of the Marshall site (as far north as the ash basin) based on average groundwater withdrawal rates. I used the CSA bedrock water -level data and all of the above information and analyses to re -construct the bedrock hydraulic head map and generate groundwater pathlines, which are shown in Figure 2. 1 conservatively assumed a small 25-foot drawdown in the Duke Energy supply well, five feet of drawdown in the Old Church well, and no pumping in the Midway Marina well. Solid black circles represent bedrock monitoring well data. Blue circles represent inferred bedrock heads beneath Lake Norman, which I assumed were equal to 760 feet in areas located beyond the radius of influence of any of the three supply wells. I experimented with lake elevations that were a few feet different than 760 feet and found that these variations had little effect on my contour map. Within the radius of influence I subtracted the drawdowns in Figure 1 (Kb, = 2.0 feet/day curve) from the 760-foot lake level. I generated Figure 2 using the kriging interpolation option in the Tecplot data visualization software package (Tecplot, Inc., Bellevue, Washington). Kriging is an advanced geostatistical procedure that spatially interpolates between discrete measurements (e.g., monitoring well water levels) by computing a weighted average of the known values of the function (hydraulic head variation) in the vicinity of the measured value. Kriging is mathematically closely related to regression analysis and generally gives the best unbiased prediction of interpolated values (Deutsch and Journel, 1992). The contours and pathlines in Figure 2 indicate that much of the bedrock groundwater flow in the southeastern portion of the area enclosed by the ash basin compliance boundary is off -site and to the south-southeast (e.g., areas west of GWA-1 BR and south of AB-613R). In the area generally northeast of wells AB-5BR and GWA-1 BR groundwater in the bedrock aquifer flows to the east-northeast toward the unnamed stream and Lake Norman. I further examined bedrock groundwater flow directions and the hydraulic effects of off -site pumping by developing a second bedrock hydraulic head map (Figure 3) in which the Duke Energy, Old Country Church, and Midway Restaurant and Marina water supply wells were inactive. Without off -site groundwater extraction from these supply wells the bedrock groundwater flow directions in the southeastern part of the Marshall site exhibit a small additional easterly component, but overall are not significantly different than the south -southeasterly directions in this area of the site shown in Figure 2. 12 Exceedances of Groundwater Standards In this section I compare measured groundwater concentrations in shallow, deep, and bedrock groundwater samples to North Carolina 2L and IMAC standards and show the following: (i) 15 measured exceedances for several COI at multiple downgradient locations on the Compliance Boundary (CB); (ii) an additional six (6) exceedances at CB locations based on chemical transport modeling I performed; (iii) 19 of the 21 downgradient Compliance Boundary exceedances were greater than the proposed provisional background concentrations (PPBC) by HDR; (iv) 13 of the 21 CB exceedances were greater than the maximum background concentration in the same hydrogeologic unit (e.g., shallow, deep, or bedrock); (v) 22 additional exceedances were observed in wells screened in the highly -permeable fractured bedrock unit underlying the ash basin and located inside the CB; and (v) the statistical analyses of groundwater concentrations at wells MW-4 and MW-41D for purposes of defining background levels were performed incorrectly. Throughout this report I reference the ash basin compliance boundary and the Duke Energy property boundary for the Marshall site as drawn on maps developed by HDR. My reference to the "compliance boundary" is only for identification purposes and not an opinion that this boundary as drawn by HDR is accurate or legally correct. Summary of Exceedances Table 1 summarizes exceedances of 2L or IMAC standards in shallow, deep, and bedrock groundwater samples obtained from monitoring wells located: (i) on downgradient sections of the Ash Basin Compliance Boundary (CB) as drawn by HDR; (ii) bedrock wells (BR) located inside the CB; and (iii) modeled Compliance Boundary concentrations (CBM), using modeling techniques described below. The proposed provisional background concentrations (PPBC) by HDR are also listed in Table 1. A total of 15 Compliance Boundary groundwater samples exceeded North Carolina groundwater standards for these COI: antimony, boron, chromium, cobalt, iron, manganese, total dissolved solids, and vanadium. I estimated an additional six (6) exceedances at downgradient CB locations based on chemical transport modeling and measured upgradient concentrations (CBM). In addition, 22 exceedances were observed in wells screened in the highly fractured bedrock unit located inside the CB. A total of 19 of the 21 (measured plus modeled) Compliance Boundary exceedances were greater than the proposed provisional background concentrations (PPBC) by HDR. A total of 13 of the 21 CB exceedances were greater than background levels from the same hydrogeologic unit (e.g., shallow, deep, 13 or bedrock) for a particular constituent. Of the 22 bedrock exceedances, 14 were greater than PPBC background levels. Note that the iso-concentration contours in all of the CSA Section 10 figures and Figure ES-1 are not consistent, and are in many cases misleading, with regard to chemical transport mechanisms in the subsurface. For example, the iso-concentration contours in Section 10 generally closely encircle a monitoring well and infer no subsequent transport downgradient from the well location. This contouring problem is especially prevalent near downgradient sections of the Compliance Boundary (e.g., the unnamed stream to the northeast and Lake Norman). Figure ES-1 is a good example of this practice. These closed contours downgradient from the ash basin near the unnamed stream suggest that boron transport beyond the farthest downgradient line of monitoring wells does not occur. This is not the case, however, as demonstrated in the following section where boron exceedances at the Compliance Boundary are demonstrated by modeling. Further, the CAP model simulated "existing conditions" plume maps for boron and several other COI (CSA Appendix C) contain 'open contours" at these downgradient parts of the CB, which confirm constituent transport to the boundary. Modeled Compliance Boundary Exceedances I computed Compliance Boundary concentrations labeled "CBM" with footnote "e" in Table 1 (wells MW- 14S,D; AL-1S,D; AB-1S; and MW-7S) using a one-dimensional, analytical chemical transport model (van Genuchten and Alves, 1982; Equation C5) because the CB at these locations was 300-600 feet downgradient from the wells and boron is highly mobile in the subsurface. I used site -specific hydrogeologic data to determine input parameter values for groundwater pore velocity (Vp) and assumed a boron retardation factor of unity (1.0), as HDR assumed in the CAP transport model. To calculate pore velocity (Vp = K i/ne ), I used: mean hydraulic conductivity (K) values from CSA Table 11-9; average effective porosity (ne) data in CSA Table 11-8; and measured horizontal hydraulic gradients (i) from the hydraulic head maps in Section 6. 1 then calibrated the model to match observed 2015 boron concentrations at the above wells. The concentrations labeled CBM in Table 1 are the modeled 2015 concentrations downgradient from each monitoring well at the Compliance Boundary. Exceedances of Groundwater and Surface Water Standards in Seep Samples As discussed in the introduction, the ash basin at the Marshall site was constructed above Holdsclaw Creek and other tributaries that discharged to the Catawba River (currently Lake Norman). Seep samples were collected during the CSA at two locations (S-2 and MSSW001 S001) on the downgradient toe of the active ash basin dam during the CSA (CAP 1 Figure 2-2). After the CSA sample collection HDR reported that location S-2 was not a separate seep from the ash basin, but rather pooled water from the MSSW001 S001 seep. Nevertheless, HDR continued to show analytical results for both locations in the CAP 1 report (Figure 2-2), and my following discussion refers to both locations. 14 Concentrations in the two seep water samples exceeded relevant NCAC 2B, 2L and/or IMAC standards for various COI (e.g., CSA Table 7-8, CAP 1 Table 2-5, and CAP 1 Figure 2-2). Referring to my Table 1, 17 exceedances of North Carolina groundwater standards were detected in the seep samples for these COI: arsenic, barium, beryllium, boron, chromium, cobalt, lead, manganese, selenium, thallium, total dissolved solids, and vanadium. North Carolina surface water (2B) standards were exceeded for these constituents: arsenic, beryllium, chromium, copper, lead, mercury, thallium, total dissolved solids, and zinc. Surface water sample SW-6, located in the unnamed tributary to Lake Norman immediately downgradient from the active ash basin (CAP 1 Figure 2-2), was also considered by HDR to be representative of groundwater quality at the site (CAP 2, Section 3.3.3). As shown in CAP 1 Table 2-6 North Carolina 2B standards were exceeded in the July and October 2015 SW-6 samples for these COI: cobalt, iron, manganese, sulfate, and total dissolved solids. Based on the significance of these exceedances, HDR recommended in the CAP 2 report (Section 3.3.3) that further evaluation of the SW-6 surface-water/seep fluid be conducted. Statistical Analyses of Background Concentrations Appendix G of the CSA report presents statistical analyses of historical concentrations from monitoring wells MW-4 (shallow) and MW-41D (deep), which HDR described as following methods specified by the U.S. Environmental Protection Agency (EPA, 2009), in an attempt to establish background groundwater concentrations for the Marshall site. As outlined in Sections 3.2.1 and 5.5.2 of the EPA guidance document these data must be checked to ensure that they are statistically independent and exhibit no pairwise correlation. Groundwater sampling data can be non -independent (i.e., autocorrelated) if the sampling frequency is too high (i.e., time interval between sampling events is too small) compared to the chemical migration rate in the aquifer (groundwater pore velocity divided by chemical retardation factor). Section 14 of the EPA guidance presents methods for ensuring that the data for background wells are not autocorrelated, but the analyses in CSA Appendix G did not include evaluations for statistical independence. As an illustration, "slow -moving" groundwater combined with high chemical retardation (i.e., large soil - water partition coefficients, Kd), which is the case for many COI in the shallow aquifer at the Marshall site, can lead to the same general volume of the chemical plume being repeatedly sampled when the monitoring events are closely spaced. Examining shallow background well MW-4, the measure hydraulic conductivity at this location is 7.0E-5 to 3.7E-4 centimeters/sec (0.2 - 1.1 feet/day; CSA Tables 11-3 and 11-4). From CSA Table 11-8 the effective porosity at this location is about 0.25. Based on CSA Figure 6- 5 and Table 6-9 the horizontal hydraulic gradient in this area is about 0.02 feet/foot. Using these values 15 the estimated groundwater pore velocity (VP = K i/ne ) near MW-4 is on the order of 6 to 30 feet/year. (Note that the groundwater velocities listed in CSA Table 11-12 are incorrect because this table is from the Allen Steam Station CSA). Shallow pore velocities are generally factors of 40 to 200 greater downgradient of the ash basin (between the ash basin and the Compliance Boundary) due to much greater horizontal hydraulic gradients (-- 4x larger) and larger hydraulic conductivity (-- 10-50x greater) in these areas. The retardation factors, Rd, based on laboratory Kd measurements (Kd — 10 cm3/g, or greater) are on the order of 100 (or greater) for many of the COI (except conservative parameters such as sulfate and boron). Therefore, the average shallow chemical migration rate near monitoring well MW-4 (VP/Rd) is on the order of 0.06 - 0.3 feet/year for many of the non -conservative COI, assuming linear equilibrium sorption (refer to discussion below). With quarterly sampling, the chemical migration distance during the time interval between sampling rounds is less than 0.1 feet (1 inch) for several COI, which is smaller than the sand pack diameter for the monitoring wells. Therefore, based on either quarterly or annual monitoring the shallow groundwater samples from monitoring well MW-4 are basically representative of the same volume of the plume (i.e., the sandpack, depending on the well purge volume) for many COI, and any measured sample concentration changes are not due to real chemical transport effects in the aquifer. In this case, this means that the groundwater samples are non -independent and that the statistical analyses of background concentrations do not satisfy the key requirements of the analysis method. CAP Groundwater Flow Model Underestimates Potential for Off -Site Chemical Migration My discussions in this section focus on limitations of the CAP groundwater flow model. I focus specifically on model boundary conditions representing Lake Norman; the overall size of the model grid and no -flow boundary conditions on the western, southern, and northern grid boundaries; the misrepresentation of groundwater flow in the fractured bedrock aquifer; and the potential for off -site groundwater flow in relation to groundwater extraction from numerous private and public water supply wells located close to the model boundaries, but not incorporated into the flow model. Lake Norman and Southern No -Flow Boundary Conditions The CAP Parts 1 and 2 groundwater flow models effectively force all groundwater (shallow, deep, and bedrock aquifers) located beneath and downgradient from the ash basin to discharge into either Lake Norman or the unnamed stream located close to the northeastern side of the ash basin, which discharges as overland flow into the lake. The CAP 1 and 2 flow models also significantly underestimate the potential for off -site flow and chemical migration in fractured bedrock. No -flow boundary conditions 16 defined along the entire western, northern, and southern model boundaries prevent any off -site groundwater flow and chemical transport in these areas (e.g., refer to Figures 1 and 4 in Appendix C of the CAP 1 Report). The CAP 1 and 2 models assign additional no -flow boundary conditions to all downgradient bedrock cells located beneath Lake Norman, which prevents groundwater from flowing beneath the lake in the bedrock aquifer. As discussed above (e.g., Figures 2 and 3), CSA hydrogeologic data and available off -site hydraulic information clearly demonstrate that the CAP 1 and 2 models boundary conditions are incorrect. The bottom surfaces (bedrock) of the flow models are also assumed to be a no -flow boundary even though the hydraulic conductivity data and measured downward hydraulic gradients at several monitoring well clusters do not support this assumption (e.g., downward flow from deep to bedrock aquifers at well clusters AB-5D/BR and AB-6D/BR). This hydraulic representation of Lake Norman in the flow models is inaccurate for many reasons. First, the lake bottom is assumed to extend all the way through the overburden aquifer to the bedrock surface, which is not the case. Second, groundwater flow beneath and adjacent to the lake is assumed to be horizontal with zero vertical flow component. Because this boundary condition does not allow groundwater to flow vertically in areas beneath and near the lake, the CAP models do not represent actual site hydrologic conditions. Third, as represented in the CAP models, neither the lower -permeability lake bed sediments nor the smaller vertically hydraulic conductivity of underlying soils restricts the potential flow rate into or out of the lake (i.e., in the CAP models a perfect hydraulic connection exists between the aquifer and Lake Norman). HDR did not evaluate the actual degree of aquifer -lake hydraulic connection in the CSA or CAP 1 and 2. In summary, due to all of these factors the potential for site groundwater and dissolved constituents to migrate off -site as underflow beneath Lake Norman and to the south-southeast as groundwater flow cannot be evaluated with the model. The CAP model should have represented Lake Norman using a "leaky -type" boundary condition in the top model layer (McDonald and Harbaugh, 1988), and the model grid should have extended farther east and south so that the above factors could have been evaluated during model calibration and sensitivity analyses. The Electric Power Research Institute technical review committee, which included the developer of the MT3D transport code used in the CAP modeling, made the same comment (December 02, 2015 memorandum entitled "Revised Marshall Model Review", which was submitted with the CAP report). A leaky boundary condition incorporates the lake/river bed permeability and thickness, the lake/river water surface elevation, and the simulated hydraulic head in the aquifer (at the base of the lake/river bed) to dynamically specify a flux (flow rate per unit bed area) into or out of the groundwater model depending on the head difference between the lake/river and aquifer. Typically, permeability and vertical hydraulic gradient measurements for the lake bed (not collected in the CSA) and flow model calibration (three-dimensional matching of simulated and measured hydraulic head measurements in the 17 aquifer) are used to determine a best -fit estimate of lake bed conductance (permeability divided by thickness) in the model. HDR did not perform this routine analysis Limitations of No -Flow Boundary Conditions and Small Model Domain Size The limited areal extent and depth of the CAP Parts 1 and 2 flow and transport model grids prevent the use of the models as unbiased computational tools that can be used to evaluate off -site migration of coal - ash constituents. For example, the model grids should have extended farther west and incorporated groundwater extraction from off -site private -home and public water -supply wells. The western no -flow boundary in the current CAP models artificially prevents any off -site flow or transport to the west in either the bedrock or overburden aquifers. The same is true for the entire northern and southern model boundaries, as discussed above, and the artificial limitations created by the eastern Lake Norman boundary condition. In addition, the downgradient boundaries of the CAP flow and transport models do not extend to the Compliance Boundary along most of the Lake Norman boundary, which prevents their use for estimating Compliance Boundary concentrations for various remedial alternatives. The bottom boundaries of the CAP models should extend much deeper because the hydraulic conductivity of the fractured bedrock zone is similar in magnitude to the overburden soils. In the present configuration the lower boundary of the CAP 1 and 2 model grids is only about 50 feet below the bedrock surface (Figure 2 in both the CAP 1 & 2 modeling appendices). Because several bedrock wells were screened to this depth the bedrock hydraulic conductivity data collected for the CSA demonstrate that imposing an impermeable model boundary at this depth is incorrect (compare similarities of mean overburden and bedrock aquifer permeabilities in CSA Table 11-9). Off -Site Groundwater Extraction Ignored The CSA and CAP Parts 1 and 2 failed to examine the strong potential for coal -ash constituents from the Marshall site to migrate with groundwater to private and public bedrock water supply wells located immediately west, south, and north of the ash basin compliance boundary. CSA Figures 4-2 and 4-5 shows the locations of private and public water supply wells near the site. The basis of my opinion includes the following: my hydraulic analyses of pumping -induced pressure reductions in the fractured bedrock unit (Figure 1) and re-evaluation of bedrock hydraulic head variations and flow directions (Figures 2 and 3), as discussed above; hydraulic conductivity measurements for the overburden and bedrock formations; three-dimensional variations in measured hydraulic head in the bedrock and overburden units; and groundwater concentration data. As discussed throughout my report, neither the CSA nor CAP Parts 1 and 2 investigations addressed the potential for off -site migration. COI's were detected in several water supply well samples (CSA Appendix B), but the CSA report did not plot these detections on a map and did not discuss their possible relationship to the Marshall site. 18 Appendix B also did not present the well construction details (e.g., well diameter and elevation range of the well screen or open bedrock interval) so that well dilution effects and potential chemical transport pathways in the bedrock unit could be evaluated. In addition, the CSA investigations and CAP 1 modeling did not include these areas west, north, and south of the Marshall site. The CAP Part 2 flow model did include a small number (four) of residential wells located inside the undersized model domain (near the northern and western model domain boundaries), but the CAP 2 modeling report (CAP 2, Appendix B) did not show simulated hydraulic head maps with these residential wells pumping and did not provide any discussion or analyses of the potential for these wells to capture COI dissolved in groundwater. The CAP Part 2 also did not increase the model grid size to incorporate the large number of residential (89) and private (4) water supply wells located very close to the site boundary (CAP 2 Figure 3-3); fix the boundary condition problems; or correct the model input data errors I have outlined so that the flow and transport models could be used to more accurately analyze the potential for off -site chemical transport. Another important model input data error is the bedrock hydraulic conductivity, which is assumed in the CAP 1 flow model to be about a factor of three to five (3-5x) lower than the overburden aquifer in different areas (Tables 2 in CAP 1 Appendix C and CAP 2 Appendix B). In the CAP 2 flow model the bedrock permeability was further lowered to a value that is about a factor of 10 to 100 (10-100x) lower than the overburden aquifer. However, the bedrock slug test results show that the mean bedrock permeability is approximately the same as the overburden permeability (CSA Table 11-9). Therefore, the CAP 1 and 2 flow models significantly restrict (incorrectly) groundwater from flowing from the overburden aquifer into the fractured bedrock unit, which causes the CAP transport models to underestimate the potential for off - site chemical migration. Ash Basin Hydraulic Boundary Condition and Groundwater Recharge Rates Based on the model description in CAP 1 Appendix C [text and Figures 4 (boundary conditions) and 5 (groundwater recharge zones)] and CAP 2 Appendix B (Figure 5) the CAP 1 and 2 flow model representations of the active ash basin are incorrect and significantly underestimate the recharge of ash basin fluids into the underlying aquifer system (e.g., refer to my preceding discussion regarding groundwater flow in bedrock and Figures 2 and 3). As noted in the CSA report (Section 2.6), the full operating pond elevation (i.e., impounded fluid level) for the active ash basin is approximately 790 feet, which is 30 feet greater than normal water level in Lake Norman. This very large hydraulic head increase in the ash basin relative to the normal static conditions that existed in the underlying aquifer system prior to construction creates very large downward groundwater flow components beneath areas of ponded ash -basin fluid and causes the near -radial horizontal flow patterns observed in the bedrock (Figures 2 and 3) and overburden (CSA Figures 6-5 and 6-6) hydraulic head contour maps. The CAP 1 and 2 flow 19 models should have used constant -head or leaky -type boundary conditions in the top layer to represent these areas of ponded fluid located inside the ash basin (e.g., McDonald and Harbaugh, 1988). The CAP Part 2 flow model underestimates leachate discharge from the active ash basin by as much as a factor of 34 in areas of ponded surface water (e.g., refer to CSA Figure 6-2). The CAP 1 model underestimates active basin leakage by as much as a factor of 93. As shown in Figure 5 of CAP 1, Appendix C, the CAP 1 flow model assumes a constant groundwater recharge rate (i.e., leakage rate) equal to 4.5 inches/year in the active ash basin. In the CAP 2 flow model the active basin leakage rate is assumed to be 12.3 inches/year (Figure 5 of CAP 2, Appendix B). However, CSA Figure 8-4.2 (cross- section B-B) shows that the vertical hydraulic gradient through the coal ash in the downgradient portion of the active ash basin is on the order of 0.5. Using Darcy's law and the mean vertical coal -ash permeability of 6.7E-5 cm/sec in CSA Table 11-10, the approximate vertical leakage rate out of the active basin is about 420 inches/year near Lake Norman (i.e., — 34 times greater than the specified CAP 2 recharge rate of 12.3 inches/year; and — 93 times greater than the specified CAP 1 recharge rate of 4.5 inches/year). Three related impacts of this incorrect active basin boundary condition are that the CAP models significantly underestimate: (i) vertical groundwater flow rates (by up to a factor of 100) through coal -ash source material in the vicinity of the downgradient portion of the active ash basin; (ii) horizontal groundwater flow and chemical transport rates downgradient from the active ash basin (by at least a factor of three; compare simulated hydraulic gradients near Lake Norman in Figure 13 of CAP 2 Appendix B with measured gradients in Figure 2-2 of the CAP 2 report); and (iii) vertical flow rates from the overburden aquifer into the fractured bedrock unit beneath ponded areas. This incorrect boundary condition representation of the active ash basin also causes the CAP models to significantly underestimate (by at least a factor of three) both the mass loading of COI into the Lake Norman and the corresponding Lake Norman surface water concentrations (attributable to coal ash ponds) that HDR estimated with their mixing model (e.g., CAP 2 report Table 4-2 and Appendix D). In addition, a site -specific distribution of groundwater recharge values should have been developed for this and the other simulation scenarios to take into account site -specific topography and soil types (e.g., runoff estimation) and climate data (precipitation, evapotranspiration, etc.; e.g., using the U.S. Army Corps of Engineers HELP Model; Schroeder et al., 1994). The predicted water table lowering due to capping, discussed below, is very sensitive to the model recharge value; therefore, HDR should have made more effort to develop a site -specific recharge -rate distribution. 20 CAP Chemical Transport Modeling Due to model calibration, model construction, and boundary -condition and input -data errors the CAP models significantly underestimate remediation time frames. As discussed in this section, reasons for this include significant underestimation of the chemical mass sorbed to soil, failure to account for slow chemical desorption rates, inaccurate analyses of water -table lowering due to capping, and flaws in the transport model calibration. Soil -Water Partition Coefficients and Model Calibration The fraction of chemical mass sorbed to soil can be represented by the soil -water partition coefficient, Kd (Lyman et al., 1982). Kd is an especially important parameter at the Marshall site because for most of the COI the bulk of the chemical mass in the soil is associated with the solid phase (i.e., sorbed to soil grains rather than dissolved in pore water). In effect, the solid fraction of the soil matrix acts as a large "storage reservoir" for chemical mass when Kd is large [e.g., metals, many chlorinated solvents, and highly - chlorinated polycyclic aromatic hydrocarbon (PAH) compounds associated with coal tars and wood - treating fluids]. Kd is also a very important chemical transport parameter which is used to compute the chemical retardation factor, Rd, assuming linear equilibrium partitioning of mass between the soil (solid) and pore -water phases (Hemond and Fechner, 1994): Rd = 1 + PbKd l ne where Pb is the soil matrix bulk dry density and ne is the effective soil porosity. For example, the chemical migration rate is directly proportional to hydraulic conductivity and inversely proportional to Rd . The total contaminant mass in an aquifer is also directly proportional to Rd, as well as aquifer cleanup times once the source is removed (e.g., Zheng et al., 1991). Accordingly, it is very important to use accurate Kd values in the CAP Closure Scenario modeling. Specifically, the CAP Part 1 transport modeling used Kd values that were typically factors of 10 - 100 (i.e., one to two orders of magnitude) smaller than the measured site -specific Kd's reported in CAP Appendix D. In contrast, the CAP Part 2 transport modeling used Kd values that are generally a factor of about 10 larger than the CAP 1 values (480x larger for arsenic; 70x larger for thallium; 25x larger for vanadium); however, the CAP 2 Kd's remain on the order of 10 times smaller than the measured site - specific Kd's reported in CAP 1 Appendix D and CAP 2 Appendix C for many COI. Further, soil -water partition coefficients for the CAP Parts 1 and 2 models are much smaller than most values presented in the literature for the COI (e.g., EPRI, 1984; Baes and Sharp, 1983). This means that, using the actual measured Kd's for the Marshall site, the times required to reach North Carolina water quality standards at 21 the Compliance Boundary are more than a factor of 10 longer (see additional discussion below) than cleanup times predicted by the CAP transport model. The CAP Part 1 modeling report (CAP 1 Appendix C; Section 4.8) argues that the major Kd reductions were needed due to the following: "The conceptual transport model specifies that COls enter the model from the shallow saturated source zone in the ash basin and beneath the CCR waste management units. When the measured Kd values were applied in the numerical model to arsenic migrating from the source zones, this COI did not reach the downgradient observation wells where it was observed in July 2015 at the end of the model simulation period. The most appropriate method to calibrate the transport model in this case is to lower the Kd values until an acceptable agreement between measured and modeled concentrations is achieved. Thus, an effective Kd value results that likely represents the combined result of intermittent activities over the service life of the ash basin, landfills, and structural fills. These may include pond dredging, dewatering for dike construction, and ash grading and placement." Considering the approach that HDR used to develop the chemical transport model (history matching), it is not true that "the most appropriate method to calibrate the transport model is to lower the Kd values." The CAP transport model used an incorrect value (2.12 g/cm3) for the bulk density of overburden materials. The bulk density should have been computed using the total porosity (n) values in CSA Tables 11-1 and 11-7 using the following formula (e.g., Baes and Sharp, 1983): Pb = 2.65 (1— n) Based on the Table 11-7 mean total porosity values (--44%) Pb — 1.48 g/cm3 for overburden soil, which means that the Rd values for the CAP 1 and 2 models were as much as a factor of 1.4 (2.12/1.48) too high before HDR adjusted the Kd values during calibration. Also, as discussed earlier, the overburden slug test values were about 50 percent too low due to HDR's data analysis errors. Both of these errors (sorption rate and hydraulic conductivity) resulted in a modeled transport rate that was as much as a factor of 2.1 too low before calibration simply due to data input errors. At least two other important factors were not considered during the CAP 1 and 2 transport model calibrations. First, the groundwater flow models are based on average hydraulic conductivity (K) values within a material zone, but K distributions in aquifers are highly variable due to layering and other types of heterogeneities (e.g., varying by factors of 3-10, or more, over distances as small as a few feet: Gelhar, 1984, 1986, 1987; Gelhar and Axness, 1983; Rehfeldt et al., 1992; Rehfeldt and Gelhar, 1992; Molz, 2015). The Marshall site hydrogeology certainly qualifies as "heterogeneous". This is very important to consider for the CAP transport model calibrations because it is the high -permeability zones and/or layers that control the time required (Tt,,,, l ) for a constituent to reach a downgradient observation point, and 22 HDR used differences in observed versus simulated Tt,,,,e, (i.e., time to travel from sources zones to downgradient monitoring wells) as the justification for lowering measured Kd values. Second, the history matching that HDR performed is very sensitive to the assumed time at which the source (i.e., coal ash) is "turned on" and to the assumed distribution of source concentrations (fixed pore water concentrations) in source area cells. Section 5.2 of CAP 1 Appendix C explains that different fixed source concentration boundary conditions were specified for the Ash Basin (1965; model year 0), Dry Ash Landfill Phases 1 & 2 (1984; model year 19), and the PV Structural Fill (2000; model year 35). These different source concentrations were assumed to remain constant following their activation in the model. For several reasons this source -area boundary condition approach is a major simplification and generally inaccurate: coal ash was gradually and nonuniformly disposed of and distributed (spatially and temporally) throughout the simulation period; it is very difficult (or not possible) to accurately extrapolate geochemical or ash -water leaching conditions (i.e., predict COI pore -water concentrations) that existed during the 2015 sampling round to conditions that may have existed in 1965 and thereafter; the actual source -area concentration distributions are highly nonuniform, but it is not clear from the CAP modeling reports how "... source concentrations were adjusted to match measured values... ", or if the source area concentrations were nonuniform. All of these uncertainties are further magnified when using history matching to calibrate a chemical transport model. Based on the above model input errors and major uncertainties in hydraulic -conductivity variations and source -term modeling, it is incorrect for HDR to simply reduce Kd values by factors of 10 to 100 below site measurements (and the large database of literature Kd values) based only on the transport model "history matching" exercises that HDR performed. My additional comments on the CAP Parts 1 and 2 transport modeling of Closure Scenarios are listed in the following section. Geochemical Modeling and Evaluation of Monitored Natural Attenuation The CAP Part 2 geochemical modeling results do not include quantitative analyses of COI attenuation rates at the Marshall site and are only qualitative in nature. In addition, HDR did not incorporate any source/sink (e.g., precipitation/dissolution) terms representing geochemical reaction mechanisms in the CAP 2 chemical transport model to evaluate whether such reactions are important compared to groundwater concentration changes caused by advection, dispersion, and soil -water partitioning. In this regard, HDR states in Section 2.10 of CAP 2 Appendix B : "A physical -type modeling approach was used, as site -specific geochemical conditions are not understood or characterized at the scale and extent required for inclusion in the model." Indeed, the Electric Power Research Institute (e.g., EPRI, 1984; page S-8) has extensively reviewed subsurface chemical attenuation mechanisms applicable to the "utility waste environment" and concluded: (i) precipitation/dissolution has not been adequately studied; and (ii) "Quantitative predictions of chemical attenuation rates based upon mineralogy and groundwater 23 composition cannot be made because only descriptive and qualitative information are available for adsorption/desorption mechanisms." Nonetheless, HDR performed geochemical modeling to evaluate the technical basis for its monitored natural attenuation (MNA) analysis; however, any quantitative MNA analysis must compare mass transport rates and changes (e.g., grams/year per unit area normal to a groundwater pathline) in the aquifer for the various active transport mechanisms in order to determine whether MNA is a viable alternative (e.g., produces meaningful groundwater concentration reductions) at the Marshall site. In Section 6.3.2 of the CAP 2 report HDR acknowledges that these quantitative evaluations were not performed in CAP 2 and indicated that they would need to be completed as part of a Tier III MNA assessment. Nevertheless, HDR suggested in the CAP 2 report that COI concentrations "will" or "may" attenuate over time without completing the necessary evaluations to reach these conclusions. HDR also states in CAP 2 Section 6.3.4 that ".... available assessment results indicate it is feasible that MNA can be used partially or entirely to remediate groundwater at the MSS site .... ". HDR claims in CAP 2, Section 6.3.3 (page 47) that "The groundwater model did not allow for removal of COI via coprecipitation with iron oxides, which likely resulted in a conservative prediction of COI transport." Finally, HDR concludes in Section 7.2.2.1 of the CAP 2 report the following regarding the MNA assessment and Appendix H: "The most significant finding was that the precipitation of iron and manganese serves to remove other COls through co -precipitation and adsorption, thus confirming that attenuation is occurring." I saw no quantitative analyses or evidence in the CAP 2 report or related appendices to support these claims. In fact, the CAP 2 Appendix H emphasizes that much more geochemical data need to be collected and chemical transport modeling with a source/sink term must be performed in a Tier III assessment to further assess whether MNA is a viable remedial alternative. Therefore, the CAP 2 report fails to provide any quantitative evidence supporting COI attenuation due to co -precipitation with iron or manganese. The second component of COI attenuation evaluated in Appendix H is chemical sorption to soil. It is important to note that, although the CAP models did not incorporate a mechanism for co -precipitation with iron or manganese (or any COI sink term), the CAP models did simulate attenuation due to sorption. Even with the sorption attenuation mechanism included, CAP 2 Table 4-1 shows that for both the "existing conditions" and "cap -in -place" scenarios the following COI will exceed North Carolina groundwater standards at the Compliance Boundary 100 years into the future: antimony, beryllium, boron, chromium, cobalt, hexavalent chromium, thallium, and vanadium. Further, my Table 1 shows that groundwater standards are currently exceeded at the Compliance Boundary for iron, manganese, and total dissolved solids (i.e., iron, manganese, and TDS contaminant plumes originating in the source areas have already reached the Compliance Boundary). The conclusions of the MNA Tier I analyses (CAP 2 Appendix H, page 17) were that antimony, beryllium, boron, cobalt, manganese, and selenium show limited evidence of attenuation and should not be 24 evaluated further for MNA. These Appendix H conclusions are consistent with the Table 4-1 results for several of these COI (antimony, beryllium, boron, and cobalt). In addition, my Table 1 results support the Appendix H conclusions that antimony, boron, cobalt, and manganese are not candidates for MNA because these contaminant plumes have also already reached the Compliance Boundary. The Table 1 results contradict the Appendix H conclusion that chromium and vanadium should be further evaluated because these COI currently exceed groundwater standards at the Compliance Boundary. All of these data and CAP 2 modeling results strongly contradict the CAP 2 conclusion (e.g., Section 7.2.2.1) that ".... MNA is the recommended corrective action for the Marshall site." Simulation of Closure Scenarios As discussed below, CAP 1 Closure Scenario simulations greatly underestimate (by factors of 10 or more) the time frames required to achieve meaningful groundwater concentration reductions in response to remedial actions. Compared to the Cap -in -Place remedial alternative evaluated in the CAP Part 1, the Excavation Scenario results in COI concentrations in groundwater that are in many cases at least a factor of 10 smaller. Although the CAP 1 modeling showed that Source Excavation outperforms CIP, the CAP 2 modeling did not simulate an Excavation closure scenario. Nonetheless, the following comparisons between CIP and Excavation impacts on groundwater concentrations are valid for both the CAP 1 and 2 model results. This is because the main difference with the CAP 2 transport model (compared to CAP 1) is that concentration changes resulting from either CIP or Excavation (if it was evaluated in CAP 2) occur much more slowly (i.e., — 10x slower; 25-480x slower for arsenic, thallium, and vanadium) in the CAP 2 model due to the much larger Kd (and Rd). The CAP 2 transport model also assumed uniform initial COI concentrations equal to HDR's proposed provisional background concentrations (PPBC), even though the PPBC exaggerate background levels (see above discussion) and there are no data to suggest that background concentrations should be spatially uniform. Despite these changes in the CAP 1 and 2 models, the relative differences in groundwater concentrations between the two closure scenarios remain about the same if the uniform starting (PPBC) COI concentrations are subtracted from the simulated concentration versus time curves. For these reasons the following discussions focus on the CAP 1 modeling results. Source Concentrations for Cap -in -Place Scenario In this scenario the CAP 1 flow model predicts limited cap -induced water -table declines: approximately 14 feet (10 feet in the CAP 2 flow modeling) beneath the downgradient portion of the ash basin; and about 2 feet, 3 feet, and 7 feet (about 1 foot in CAP 2) beneath the dry ash landfill (Phase 1), Phase II of the dry ash landfill, and the PV structural fill, respectively. The geologic cross -sections presented in the CSA show that the saturated coal ash thickness in the ash basin generally varies from 25 to 70 feet and 25 is on the order of 50 feet at many locations. Therefore, these cap -induced water table declines are about a factor of four too small at most locations in the ash basin to dewater the source material. For example, the CSA geologic cross -sections show saturated coal ash thicknesses in the active ash basin equal to about 40-50 feet in cross-section A -A (northwest to southeast section generally through the middle of the ash basin); 50-70 feet in cross-section B-B (southwest portion of the ash basin); and 25-45 feet in cross- section C-C (north-northeast portion of the ash basin). This means that under the simulated Cap -In -Place Scenario at the majority of locations in the ash basin at least 75 percent of the coal ash, which is the source of dissolved COI, would remain saturated and continue to leach constituents into groundwater. The CAP 1 simulations ignored this fact and set all source concentrations equal to zero (i.e., assumed all coal ash was dewatered). Therefore, the simulated Cap -in -Place concentrations should be much higher than the values presented in the CAP Part 1. It is important to note, however, that the CAP Parts 1 and 2 groundwater flow model simulations exaggerate the hydraulic effects of the cap (i.e., overstates water table lowering) due to the lateral no -flow boundaries and the no -flow boundary at the base of the model, which is located only 50 feet below the top of the highly -permeable, fractured bedrock aquifer (— 100 feet below ground surface). Note that the depths of the neighboring Duke Energy and Midway Marina bedrock water supply wells, discussed above, are 540 feet and 245 feet below ground surface (about 400 feet and 160 feet below the bedrock surface), respectively, and provide high groundwater extraction yields of 30 and 20 gallons per minute, respectively [Table 1 of the 2014 receptor survey report (HDR, 2014)]. The no -flow boundary conditions along the entire western, southern, and northern grid boundaries prevent flow into the ash basin when large, laterally inward hydraulic gradients are created by capping. In addition, the no -flow boundary condition at the base of the flow model prevents upward flow from highly -permeable portions of the bedrock aquifer that underlie the base of the CAP models, thus further exaggerating predicted water table lowering. Slow and Multirate Nonequilibrium Desorption of COI Since the 1980's the groundwater industry has learned how difficult it is to achieve water quality standards at remediation sites without using robust corrective actions such as source removal (Hadley and Newell, 2012, 2014; Siegel, 2014). Two of the key reasons for this in aqueous -phase contaminated soil are inherently low groundwater or remediation fluid flushing rates in low -permeability zones and slow, nonequilibrium chemical desorption from the soil matrix (Culver et al., 1997, 2000; Zheng et al., 2010). A good example of this is the "tailing effect" (i.e., very slow concentration reduction with time) that is commonly observed with pump -and -treat, hydraulic containment systems. These factors are also related to the "rebound effect" in which groundwater concentrations sometimes increase shortly after a 26 remediation system is turned off (Sudicky and Illman, 2011; Hadley and Newell, 2014; Culver et al., 1997). The CAP 1 and 2 flow models use different permeability (K) zones, but the scale of these zones is very large and within each zone K is homogeneous even though large hydraulic conductivity variations (e.g., lognormal distribution) are known to exist at any field site over relatively small length scales (Molz, 2015). Moreover, the CAP transport models assume linear, equilibrium soil -water partitioning which corresponds to instantaneous COI release into flowing groundwater. The transport code (MT3D) has the capability of simulating single -rate nonequilibrium sorption, but the Close Scenario simulations did not utilize this modeling feature. Slow desorption of COI can also be expected at the Marshall site because sorption rates are generally highly variable and multi -rate (Culver et al., 1997, 2000; Zheng et al., 2010), and Kd values are nonuniform spatially (Baes and Sharp, 1983; EPRI, 1984; De Wit et al., 1995). The CAP flow and transport models can be expected to significantly underestimate cleanup times required to meet groundwater standards at the compliance boundary because they do not incorporate these important physical mechanisms. Adequacy of the Kd Model for Transport Simulation The laboratory column experiment effluent data (e.g., CAP 1 Appendix D) generally gave very poor matches with the analytical (one-dimensional) transport model used to compute Kd values. Since the CAP transport models solve the same governing equations in three dimensions, the adequacy of the Kd modeling approach for long-term remedial simulations should have been evaluated in much more detail in the modeling appendix. The transport modeling also did not evaluate alternative nonlinear sorption models such as the Freundlich and Langmuir isotherms (Hemond and Fechner, 1994), which are input options in the MT31D transport code. Several of the batch equilibrium sorption experiments (CAP 1 Appendix D) exhibited nonlinear behavior, and such behavior is commonly observed in other studies (e.g., EPRI, 1984). However, HDR only computed linear sorption coefficients (i.e., Kd) for the Marshall site in CAP Part 1. In CAP Part 2 HDR did fit Freundlich isotherms to the batch sorption data for selected COI (CAP 2 Appendix C, Tables 1-8) but did not use these Freundlich isotherm results in the CAP 2 transport modeling. De Wit et al. (1995) showed that the nonlinear sorption mechanism is similar in importance to aquifer heterogeneities in extending remediation time frames. Closure Scenario Time Frames Many of the simulated concentration versus time (CvT) curves presented in CAP 1 Appendix C are either incorrect or highly misleading with respect to future groundwater concentration variations that may occur under the Existing Conditions, Cap -In -Place (CIP), and Excavation scenarios. For example, several CvT 27 curves suggest that concentrations of several COI at various locations in the aquifer system would be as much as factors of 10 to 100 greater if all of the ash source material is removed (Excavation) compared to either doing nothing (Existing Conditions) or CIP for a period of 200 years into the future (e.g., arsenic, antimony, boron, chloride, cobalt, chromium, hexavalent chromium, selenium, sulfate, thallium, and vanadium at AB-1 BR; antimony at AB-5BR; arsenic, selenium, and vanadium at AB-6BR; barium at AL- 2D; beryllium at AB-1 S & AL-1 S; chromium at AB-3D; and selenium, sulfate, and vanadium at AL-4D). Further, the model results for some of these CvT curves show groundwater concentrations increasing at a dramatic rate 200 years into the future even if all source material is removed (e.g., selenium and vanadium at AB-6BR; arsenic, cobalt, and hexavalent chromium at AB-1 BR; arsenic at AB-613R). The CvT curves are also highly inconsistent at the same location (e.g., for the Excavation Scenario at AB-1 BR arsenic concentrations increase significantly for 200 years, but boron concentrations decrease significantly for the same time period) and between different locations (e.g., for the same COI the CvT curves suggest that at some locations Excavation is much better than CIP, but at other locations CIP is much better than Excavation). In addition to these issues, the CAP 1 chemical transport model underestimates the time intervals required to achieve groundwater concentration reductions (i.e., achieve groundwater quality restoration) by a factor of 10 or more, as discussed in different parts of my report. In other words, the CAP 1 transport model significantly overestimates the rate at which concentrations may reduce in response to remedial actions such as capping or source removal. This is due to several factors, including major errors in model input data, model calibration mistakes, field data analysis errors, and oversimplified model representation of field conditions (e.g., hydraulic conductivity) and transport mechanisms (e.g., chemical sorption/desorption). These limitations of transport models for realistically predicting cleanup times have been recognized by the groundwater industry for the past few decades based on hands-on experience at hundreds of extensively -monitored remediation sites. Even if we ignore the above problems with the CvT curves in CAP 1 Appendix C and the factors of 10 or more errors in cleanup time predictions with the CAP model, the remediation time frames for the Excavation Scenarios are still more than two centuries for several constituents due to slow groundwater flushing rates from secondary sources (surrounding residual soil) left in place after excavation and due to high chemical retardation factors for most of the COI. However, excavation of secondary -source material would further accelerate cleanup rates under this alternative. The simulated Cap -In -Place concentration reduction rates are much slower, compared to excavation, but are also incorrect (i.e., overestimated) because the cap -induced water -table lowering only dewaters a small percentage (— 25 percent) of the source -area coal ash, as discussed above, and the CAP flow models overestimate cap -induced water - table lowering due to boundary condition errors. Furthermore, these simulation time frames are well beyond the prediction capabilities of any chemical transport model for a complex field site, especially one 28 that is as geochemically complex as the Marshall site. The historical model -calibration dataset (1965- 2015) is also significantly smaller than the predictive (remediation) time frames. In addition, the "history matching" technique used to calibrate the transport model (e.g., major reduction in measured Kd values) was not performed correctly by HDR. Cap -In -Place versus Excavation Closure Scenarios Even though the CAP 1 model significantly underestimates remediation time frames, the CAP 1 Closure Scenario simulations (using CvT curves that do not appear to be suspect) demonstrate several significant advantages of excavation for restoring site groundwater quality versus CIP. First, predicted COI concentration reductions in groundwater are in many cases at least a factor of 10 greater with excavation compared to CIP [e.g., antimony and boron (AB-31D, AB-5BR); arsenic, barium, cobalt, chromium, hexavalent chromium, thallium (AB-3D); beryllium, cobalt, chromium, sulfate, thallium (AB-5BR); refer to CAP Appendix C]. Further, if the cap -in -place simulations would have been performed correctly the simulated cap -in -place concentrations would be much higher because predicted water -table lowering due to the cap would be insufficient to dewater all of the coal ash. Although the CAP 1 modeling demonstrated that the CIP closure alternative would be much less effective than excavation, and that CIP would only dewater about one-fourth of the saturated coal -ash thickness in many areas, HDR eliminated excavation from consideration in CAP 2. In Section 7.1 of the CAP 2 report HDR assumes that "groundwater flow and geochemical modeling indicates that attenuation by a combination of sorption, chemical precipitation, and dilution by surface water infiltration and fresh groundwater effectively dissipates COls in groundwater beneath and downgradient of the source areas" and that "... it is reasonable to assume that COls remaining in groundwater will decrease in concentration overtime as upgradient non -impacted water moves through the aquifer." As discussed above, HDR provided no quantitative analysis or evidence in the CAP 2 report or related appendices to support this claim. Considering that up to 75 percent of the coal -ash source material would remain saturated with CIP and that multiple exceedances of groundwater standards at the Compliance Boundary currently exist (with no historical data to indicate that these Compliance Boundary concentrations are decreasing with time), it is not reasonable to make sweeping assumptions about future concentration changes. Tier III MNA analyses require rigorous quantitative evaluations using the CAP transport model with a source/sink term that incorporates geochemical reactions to support MNA as a viable corrective action. CAP 2 did not provide this information. As discussed above, the CAP Part 2 flow model did include a very small number of residential wells (4 of the 89 neighboring private wells and none of the four public water supply wells), but the CAP 2 modeling report (CAP 2, Appendix B) did not show simulated hydraulic head maps with these residential wells pumping and did not provide any discussion or analyses of the long-term potential for these wells to 29 capture COI dissolved in groundwater. Further, the private bedrock wells that HDR chose to include in the CAP 2 model are located upgradient from the ash basins; HDR should have included all of the private wells located immediately north, west, and south of the Duke Energy property boundary and the four public water supply wells to the south (refer to my previous discussion revised maps of bedrock groundwater flow directions). In CAP 2 section 4.1.5 HDR discusses that fact that the CAP 2 flow model was used to compute 1-year, reverse particle pathlines for these bedrock residential wells (Figure 15 in CAP 2 Appendix B) to determine their short-term groundwater capture zones. However, the residential well reverse pathline tracing should have been performed for a much longer time period (e.g., from the beginning of coal ash disposal to the present) to evaluate whether COI may have migrated from source areas to these wells. In addition, if HDR had extended the CAP 2 model grid much farther to the north, west, and south the capture zones for the remaining 85 private and four public water supply wells could have been determined, as I discuss earlier in my report. The CAP Closure Scenarios do not include hydraulic containment remedial alternatives (e.g., gradient reversal) for the bedrock aquifer that would address the risk of off -site COI transport. As discussed above, the CSA data show many exceedances of groundwater standards in bedrock inside the compliance boundary. In addition, downward groundwater flow components from the deep overburden to bedrock aquifers were measured during the CSA in the southeastern portion of the site where groundwater flow in the bedrock aquifer is off -site (Figures 2 and 3), and ponded water in the ash basin impoundment has historically created a large downward hydraulic gradient beneath the ash basin. The cap -in -place alternative does not address either concentration reduction or off -site chemical migration control in the fractured bedrock aquifer. Conclusions Based on my technical review and analyses of the referenced information for the Marshall site I have reached the following conclusions: • A total of 21 Compliance Boundary groundwater samples exceeded North Carolina groundwater standards for these COI: antimony, boron, chromium, cobalt, iron, manganese, total dissolved solids, and vanadium. Of these 21 exceedances, 19 were greater than the proposed provisional background concentrations by HDR, which exaggerate background levels; • The statistical analyses of historical, shallow -aquifer background groundwater concentrations at the Marshall site (monitoring well MW-4) are invalid. The time periods between groundwater sample collection from this well are too small and the concentration data are not independent; 30 • There is a significant risk of chemical migration from the ash basin to neighboring private and public water supply wells in fractured bedrock. The design of the CAP Parts 1 and 2 flow and transport models prevents the potential for off -site migration from being evaluated; • The limited CAP model domain size (Parts 1 and 2); the no -flow boundary conditions along the western, southern, and northern boundaries; the impermeable boundary at the base of the model at a depth of only about 50 feet below the top of the fractured bedrock aquifer; the incorrect hydraulic boundary condition representations of Lake Norman and the active ash basin; and HDR's failure to include bedrock -aquifer groundwater extraction from most neighboring private and public water supply wells prevent simulation and analysis of off -site COI migration; • The CAP Closure Scenario simulations greatly underestimate (by factors of 10 or more) the time frames required to achieve meaningful groundwater concentration reductions in response to remedial actions. This is due to oversimplification of field fate and transport mechanisms in the CAP Parts 1 and 2 models and several model input errors; • The simulated water table lowering for the Cap -in -Place Scenario is about a factor of four too small at most locations in the ash basin in order to dewater all source material. In addition, the actual cap -induced water table elevation reduction would be much less than predicted due to the incorrect no -flow boundary conditions (3 of 4 lateral boundaries and the bottom bedrock surface) in the flow model. Therefore, the remediation time frames for this scenario would be much greater because most of the source zone would still be active with the cap installed; • For either the Existing Condition or Cap -in -Place Model Scenario groundwater concentrations of many coal -ash constituents much higher than background levels will continue to exceed North Carolina groundwater standards at the Compliance Boundary because saturated coal -ash material and secondary sources will remain in place; • Source -area mass removal included in the Excavation Scenario results in COI groundwater concentration reductions that are at least a factor of 10 greater compared to Cap -in -Place at many locations. Additional excavation of secondary sources would further accelerate concentration reductions; • Due to an incorrect boundary -condition representation of the active ash basin, the CAP models underestimate by a factor of three or more both the mass loading of COI into Lake Norman and the corresponding Lake Norman water concentrations (attributable to coal ash ponds) estimated by the groundwater/surface-water mixing model; • The CAP Part 2 geochemical modeling and monitored natural attenuation (MNA) evaluations do not provide the required quantitative analyses (e.g., numerical transport modeling) of COI attenuation rates necessary to support MNA as a viable corrective action and are only qualitative in nature. The CAP 2 chemical transport modeling, which included attenuation by sorption, 31 demonstrated that MNA is not an effective remedial option for several COI (e.g., antimony, beryllium, boron, chromium, cobalt, hexavalent chromium, thallium, and vanadium); • The CAP Closure Scenarios do not include hydraulic containment remedial alternatives for the bedrock aquifer and do not address the risk of off -site COI transport. CSA data show 22 exceedances of groundwater standards in bedrock inside the Compliance Boundary. The cap -in - place alternative does not address off -site chemical migration control in the fractured bedrock aquifer; and • Future Compliance Monitoring at the site should include much more closely -spaced Compliance Wells to provide more accurate detection, and the time intervals between sample collection should be large enough to ensure that the groundwater sample data are statistically independent to allow accurate interpretation of concentration trends. 32 References Baes, C.F., and R.D. Sharp. 1983. A Proposal for Estimation of Soil Leaching and Leaching Constants for Use in Assessment Models. Journal of Environmental Quality, Vol. 12, No. 1. 17-28. Barker, J.A., and J.H. Black. 1983. Slug Tests in Fissured Aquifers. Water Resources Research. Vol. 19, No. 6. 1558-1564. Bear, J. 1979. Hydraulics of Groundwater. New York: McGraw-Hill. Bouwer, H., and R.C. Rice. 1976. A Slug Test for Determining Hydraulic Conductivity of Unconfined Aquifers with Completely or Partially Penetrating Wells. Water Resources Research. Vol. 12, No. 3. 423-428. Culver, T.B., S.P. Hallisey, D. Sahoo, J.J. Deitsch, and J.A. Smith. 1997. Modeling the Desorption of Organic Contaminants from Long -Term Contaminated Soil Using Distributed Mass Transfer Rates. Environmental Science and Technology, 31(6), 1581-1588. Culver, T.B., R.A. Brown, and J.A. Smith. 2000. Rate -Limited Sorption and Desorption of 1,2- Dichlorobenzene to a Natural Sand Soil Column. Environmental Science and Technology, 34(12), 2446-2452. Daniel, C.C., D.G. Smith, and J.L. Eimers. 1989. Chapter C, Hydrogeology and Simulation of Ground - Water Flow in the Thick Regolith-Fractured Crystalline Rock Aquifer System of Indian Creek Basin, North Carolina. U.S. Geological Survey Water -Supply Paper 2341-C. Ground -Water Resources of the Piedmont -Blue Ridge Provinces of North Carolina. Deutsch, C.V., and A.G. Journel. 1992. GSLIB, Geostatistical Software Library and User's Guide. Oxford University Press. De Wit, J.C.M., J.P. Okx, and J. Boode. 1995. Effect of Nonlinear Sorption and Random Spatial Variability of Sorption Parameters on Groundwater Remediation by Soil Flushing. Groundwater Quality: Remediation and Protection, Proceedings of the Prague Conference, May 1995. IAHS Publication No. 225. EPA. 1985. Full -Scale Field Evaluation of Waste Disposal from Coal -Fired Electric Generating Plants. Report EPA-600/7-85-028a, June 1985, Volume I, Section 5. Prepared by Arthur D. Little, Inc. EPA. 2009. Statistical Analysis of Groundwater Monitoring Data at RCRA Facilities - Unified Guidance. Office of Resource Conservation and Recovery, Program Implementation and Information Division, U.S. Environmental Protection Agency. Report EPA 530/R-09-007. March 2009. EPRI. 1984. Chemical Attenuation Rates, Coefficients, and Constants in Leachate Migration. Volume 1: A Critical Review. Electric Power Research Institute Report EA-3356, Volume 1. Prepared by Battelle, Pacific Northwest Laboratories, Richland, Washington. February 1984. Gelhar, L.W. 1984. Stochastic analysis of flow in heterogeneous porous media. In Selected Topics in Mechanics of Fluids in Porous Media, edited by J. Bear and M.Y. Corapcioglu, pp. 673-717, Martinus Nijhoff, Dordrecht, Netherlands. Gelhar, L.W. 1986. Stochastic subsurface hydrology from theory to applications. Water Resources Research 22: 135S-145S. 33 Gelhar, L.W. 1987. Stochastic analysis of solute transport in saturated and unsaturated porous media. In Advances in Transport Phenomena in Porous Media, NATO ASI Ser., edited by J. Bear and M.Y. Corapcioglu, pp. 657-700, Martinus Nijhoff, Dordrecht, Netherlands. Gelhar, L.W., and C.L. Axness. 1983. Three-dimensional stochastic analysis of macrodispersion in aquifers. Water Resources Research 19, no. 1: 161-180. Hadley, P.W., and C.J. Newell. 2012. Groundwater Remediation: The Next 30 Years. Groundwater, Vol. 50, No. 5. 669-678. Hadley, P.W., and C.J. Newell. 2014. The New Potential for Understanding Groundwater Contaminant Transport. Groundwater, Vol. 52, No. 2. 174-186. Hantush, M.S. 1964. Hydraulics of Wells. Advances in Hydroscience. Vol. 1. Academic Press. Ed. V.T. Chow. 282-437. Hemond, H.F., and E.J. Fechner. 1994. Chemical Fate and Transport in the Environment. Academic Press. HDR. 2014. Receptor Survey, Marshall Steam Station Ash Basin, NPDES Permit NC0004987. Prepared for Duke Energy by HDR Engineering, Inc. of the Carolinas, October 10, 2014. LeGrand, H.E. 2004. A Master Conceptual Model for Hydrogeological Site Characterization in the Piedmont and Mountain Region of North Carolina. Prepared for North Carolina Department of Environment and Natural Resources, Division of Water Quality, Groundwater Section. Lyman, W.J., W.F. Reehl, and D.H. Rosenblatt. 1982. Handbook of Chemical Property Estimation Methods. McGraw-Hill Book Company. McDonald, M.G., and A.W. Harbaugh. 1988. MODFLOW, A Modular Three -Dimensional Finite - Difference Ground -Water Flow Model. Techniques of Water -Resources Investigations of the United States Geological Survey, Department of the Interior. McGary, M. 2004. Annual Water Use Report, Marshall Steam Station, 2004. North Carolina Division of Water Resources, Water Withdrawal and Transfer Registration. Molz, F. 2015. Advection, Dispersion, and Confusion. Groundwater, Vol. 53, No. 3. 348-353. Pippin, C.G., M.J. Chapman, B.A. Huffman, M.J. Heller, and M.E. Schelgel. 2008. Hydrogeologic Setting, Ground -Water Flow, and Ground -Water Quality at the Langtree Peninsula Research Station, Iredell County, North Carolina, 2000-2005. Scientific Investigations Report 2008-5055, U.S. Department of the Interior, U.S. Geological Survey. Rehfeldt, K.R., and L.W. Gelhar. 1992. Stochastic analysis of dispersion in unsteady flow in heterogeneous aquifers. Water Resources Research 28, no. 8: 2085-2099. Rehfeldt, K.R., J.M. Boggs, and L.W. Gelhar. 1992. Field study of dispersion in a heterogeneous aquifer, 3, geostatistical analysis of hydraulic conductivity. Water Resources Research 28, no. 12: 3309- 3324. Schroeder, P.R., T.S. Dozier, P.A. Zappi, B.M. McEnroe, J.W. Sjostrom, and R.L. Peyton. 1994. The Hydrologic Evaluation of Landfill Performance (HELP) Model: Engineering Documentation for Version 3. Report EPA/600/R-94/168b, September 1994, U.S. Environmental Protection Agency, Office of Research and Development, Washington, D.C. 34 Siegel, D.I. 2014. On the Effectiveness of Remediating Groundwater Contamination: Waiting for the Black Swan. Groundwater, Vol. 52, No. 4. 488-490. Simon, R.B., S. Bernard, C. Meurville, and V. Rebour. 2105. Flow -Through Stream Modeling with MODFLOW and MT3D: Certainties and Limitations. Groundwater, Vol. 53, No. 6. 967-971. Sudicky, E.A., and W.A. Illman. 2011. Lessons Learned from a Suite of CFB Borden Experiments. Groundwater, Vol. 49, No. 5. 630-648. van Genuchten, M.Th., and W.J. Alves. 1982. Analytical Solutions of the One -Dimensional Convective - Dispersive Solute Transport Equations. U.S. Department of Agriculture, Agricultural Research Service, Technical Bulletin No. 1661. Winter, T.C., D.O. Rosenberry, and J.W. La Baugh. 2003. Where Does the Ground Water in Small Watersheds Come From?. Groundwater, Vol. 41, No. 7. 989-1000. Yin, Z.-Y., and G.A. Brook. 1992. The Topographic Approach to Locating High -Yield Wells in Crystalline Rocks: Does It Work?. Groundwater, Vol. 30, No. 1. 96-102. Zheng, C., G.D. Bennett, and C.B. Andrews. 1991. Analysis of Ground -Water Remedial Alternatives at a Superfund Site. Groundwater, Vol. 29, No. 6. 838-848. Zheng, C., M. Bianchi, S.M. Gorelick. 2010. Lessons Learned from 25 Years of Research at the MADE Site. Groundwater, Vol. 49, No. 5. 649-662. 35 Figures 36 10' 10' 00 10, C 0 i 10-1 0 10" 10' KBR = 0.68 fWay ���• KBR=4.7fflday Q K,=2.0ftlday \ i i i14 • i i • i ■ r 14 ` • r "O r " r ■ • i r i 1000 2000 3000 Distance from Pumping Well (feet) Figure 1 Drawdown in Bedrock Versus Distance from Pumping Well 37 4000 BG-2BR �r i:i Sfl4 1 GWA-9BR = 1 1 --a' AB•15BR -.. Ate' i AL-2BR r P tAB-GBRJ _` i 4 p qpABABR. ` ... Lake Norman j 1 Ash Basin Waste Boundary - i, AB-5BR � e`�'. t ' r I • /' j \ 1 � F e Lake Norman ~GWA-1BR % •j• • tr �$ DId Country Church V/ Duke Supply Well •1 I TI f \ 1 `v Figure 2 Bedrock Hydraulic Head Map with Off -Site Pumping 38 GlR.I B' R� I w AB-68R Ash Basin Waste Boundary BG-26R 726 MW-14BR Ip AB-1BR Lake Norman Lake Norman BR *I Old CouELty Church] lre F0 Figure 3 Bedrock Hydraulic Head Map without Off -Site Pumping 39 Table 40 Table 1. Exceedances of NC Groundwater Standards at Compliance Boundary and in Bedrock for 2015 Monitoring Well and Seep Samples Constituent Antimony Background Sample Sample Location Concentration Concentrationsc,d Name Depth Measured PPBC GWA-2D Deep CB 4.1 AB-6BR Bedrock BR 1.3 Arsenic S-2 Seep Ash Basin 87.1 Barium S-2 Seep Ash Basin 990 Beryllium S-2 Seep Ash Basin 15.2 Boron MW-14S Shallow CB Me 2,500e AL-1 S CB Me 4,300e AB-1 S CB Me 4,500e MW-7S CB Me 4,400e MW-14D Deep CB Me 2,500e AL-1 D CB Me 1,100e AL-2BR Bedrock BR 2,100 S-2 Seep Ash Basin 4,000 MSSW001 S001 6,000 41 Standard 2L IMAC 2_5 1 - ND (BG-1 D, BG-3D), 0.33 (MW-4D) ND (BG-2BR) 5 10 - 157.3 700 - 1 4 - 100 700 - ND (BG-1S,-2S,-3S; MW-4) ND (MW-4D, BG-1 D), 26 (BG-3D) ND (BG-2BR) Table 1. Continued Constituent Sample Name Sample Depth Location Concentration' Background Concentrations,,d Standard' Measured PPBC 2L IMAC Chromium GWA-7S Shallow CB 22.1 11.3 10 - 5.7 (BG-1 S), 9 (BG-2S), 2.5 (MW-4), 73.7 (BG-3S) GWA-2D Deep CB 182 3.1 (BG-1 D), 6.6 (BG-3D), 1.2 (MW-4D) AL-2BR BR 17.5 80.4 (BG-2BR) S-2 Seep Ash Basin 85.7 Cobalt GWA-2S Shallow CB 2.6 2.5 - 1 1.2 (BG-1 S), 0.38 (BG-2S), ND (MW-4), 5 (BG-3S) AB-5BR Bedrock BR 7.9 GWA-1 BR BR 1.7 11.9 (BG-2BR) S-2 Seep Ash Basin 333 MSSW001 S001 92 Iron GWA-7S Shallow CB 12,300 467.1 300 - 480 (BG-1 S), 140 (BG-2S), 510 (MW-4), 3,400 (BG-3S) GWA-7D Deep CB 2,200 310 (BG-1 D), 250 (BG-3D), 77 (MW-4D) 42 Table 1. Continued Constituent Sample Name Sample Depth' Location Concentration` Background Concentrationsc,d Measured PPBC Standard 2L IMAC AB-1 BR Bedrock BR 2,800 467.1 300 - Iron GWA-1 BR BR 780 (continued) MW-14BR BR 320 18,200 (BG-2BR) AB-15BR BR 1,300 Lead S-2 Seep Ash Basin 227 35.9 15 - Manganese GWA-2S Shallow CB 92 48 50 - GWA-7S CB 510 160 (BG-1 S), 20 (BG-2S), 19 (MW-4), 180 (BG-3S) GWA-7D Deep CB 76 54 (BG-1 D), 24 (BG-3D), 4.1 (MW-4D) AB-1 BR Bedrock BR 54 GWA-1 BR BR 780 380 (BG-2BR) AB-5BR BR 650 AB-15BR BR 270 S-2 Seep Ash Basin 11,600 MSSW001 S001 8,500 Selenium AL-2BR Bedrock BR 24 1.6 (BG-2BR) 10 20 - S-2 Seep Ash Basin 25.1 43 Table 1. Continued Constituent Sample Name Sample Depth' Location Concentration` Background Concentrationsc,d Standard Measured PPBC 2L IMAC Thallium S-2 Seep Ash Basin 8.6 0.5 - 0.2 MSSW001 S001 0.6 Total GWA-2D Deep CB 650,000 85,400 500,000 - Dissolved 142,000 (BG-1 D), 183,000 (BG-3D), Solids 82,000 (MW-4D) S-2 Seep Ash Basin 892,000 MSSW001 S001 989,000 Vanadium GWA-2S Shallow CB 0.33 3.9 - 0.3 GWA-7S CB 23.7 3.8 (BG-1 S), 44 (BG-2S), 2.2 (MW-4), 14.4 (BG-3S) GWA-7D Deep CB 6.1 3.5 (BG-1 D), 21.9 (BG-3D), MW-10D CB 3.2 2.8 (MW-4D) GWA-2D CB 12.2 GWA-1 BR Bedrock BR 3.5 100 (BG-2BR) MW-14BR BR 4.3 AB-6BR BR 7.4 GWA-9BR BR 3.9 AB-5BR BR 0.76 AB-1 BR BR 0.43 AB-9BR BR 2.4 AL-2BR BR 4.5 S-2 Seep Ash Basin 566 44 a Refer to the CSA Report for Monitoring Well locations and screen intervals b CB = Monitoring Well located on the Ash Basin Compliance Boundary; BR = Monitoring Well screened in Bedrock `All values are total measured concentration in the water samples in units of micrograms per liter d Refer to the CSA Report for information regarding placement of Background Wells. PPBC = Proposed Provisional Background Concentration (CAP Report Table 2-2) e CBM = Projected downgradient concentration at Compliance Boundary based on calibrated one-dimensional, analytical chemical transport model discussed in this Report 45 DOUGLAS J. COSLER, Ph.D., P.E. 10240 Stonemede Lane Matthews, NC 28105 EDUCATION 704-246-7702 dcosler@adaptivegroundwater.com Ph.D. Chemical Hydrogeology The Ohio State University 2006 C.E.D. Civil Engineer Degree Massachusetts Institute of Technology 1987 M.S. Civil & Environmental Engineering The Ohio State University 1979 B.S. Civil & Environmental Engineering The Ohio State University 1977 Summa Cum Laude PROFESSIONAL HISTORY 2009- Principal Hydrogeologist and Commercial Software Developer, Adaptive Groundwater Solutions LLC, Charlotte, NC 2007-2009 Environmental Consultant, Hart Crowser, Portland, OR 2006-2007 Research Scientist and Instructor, School of Earth Sciences, The Ohio State University, Columbus, OH 2003-2006 Research Assistant, School of Earth Sciences, The Ohio State University, Columbus, OH 1987-2003 Environmental Consultant, MACTEC (now AMEC), Nashua, NH 1984-1987 Research Assistant, Department of Civil & Environmental Engineering, Massachusetts Institute of Technology, Cambridge, MA 1979-1984 Environmental Consultant, D'Appolonia Consulting Engineers, Pittsburgh, PA 1977-1979 Research Assistant, Department of Civil & Environmental Engineering, The Ohio State University, Columbus, OH REGISTRATION Registered Professional Engineer: Pennsylvania and Vermont HONORS AND AWARDS Member of Tau Beta Pi University Graduate Fellowship, The Ohio State University, 1979 The Brown Scholarship (top undergraduate in Civil Engineering), The Ohio State University, 1977 PROFESSIONAL EXPERIENCE Environmental Consulting 1979-1984, 1987-2003, 2007-present • Areas of Specialization: Groundwater flow and chemical transport analyses and computer modeling, contaminant fate and transport in the environment, numerical code development, ground water and surface water hydraulics and hydrology, contaminant fate and transport, expert witness testimony and litigation support, hydrogeologic investigation, nonaqueous phase liquid (LNAPL/DNAPL) Douglas J. Cosler, Ph.D., P.E. - Page 2 of 15 investigation, subsurface remediation and remedial design, natural attenuation and risk assessment, and hydrologic and wetlands impact evaluation. Responsibilities: Principal Hydrogeologist/Hydrologist responsible for technical aspects of a wide variety of projects, including: investigation, remediation, and regulation of Superfund, RCRA, and other hazardous waste sites; ground water flow and chemical transport model development for numerous projects; expert witness testimony and litigation support for several clients and hazardous waste sites; natural attenuation/biodegradation assessments for chlorinated solvent and petroleum contamination sites; volatile organic compound vapor (soil gas) migration and exposure assessment; exposure modeling for health risk assessments; hydraulic and hydrologic modeling of impoundments and spillways for U.S. Army Corps of Engineers dam safety assessments; stream hydraulics and solute transport modeling; hydrologic impact assessment for minerals and coal mining; leachate collection system modeling and design for waste disposal impoundments; and design of runoff, sedimentation, and erosion control plans. Types of Sites and Contaminants: Sites investigated include: landfills, manufactured gas plants, wood - treating facilities, chemical plants, water supply well fields, gasoline and fuel oil storage/delivery facilities, nuclear waste disposal sites, hazardous waste incinerators, mining operations, and various industrial facilities. Investigated dissolved, nonaqueous-phase (LNAPL/DNAPL), and vapor -phase contaminants: chlorinated solvents, gasoline and fuel oil constituents, wood -treating products (e.g., creosote and pentachlorophenol), coal tars, polychlorinated biphenyls, pesticides, dioxins and furans, phenolic compounds, flame retardants (PBDE), phthalates, radionuclides, biological constituents, and various metals. • Representative Project Experience: Expert Witness Testimony and Litigation Support Litigation and Expert Witness Support, Wells G&H Superfund Site, Woburn, MA (MACTEC). Doug provided technical support for property owners involved in litigation related to economic damages associated with groundwater contamination in a fractured bedrock aquifer resulting from upgradient sources of chlorinated solvents (DNAPL and aqueous -phase). He completed a thorough review ofRI/FS technical reports (including groundwater pumping tests) and performed modeling of chemical transport in the fractured bedrock aquifer that accounted for the effects of horizontal anisotropy on transport directions. Based on the evaluations, Doug developed an alternative site conceptual model that incorporated the effects of bedrock fractures on solute transport in order to define probable contaminant migration pathways in overburden and bedrock aquifers that were not identified in historical documents. He demonstrated the existence of these pathways using two-dimensional models of groundwater flow and contaminant advection (particle pathlines) that established a connection between DNAPL sources areas and groundwater contamination beneath the subject properties. Expert Witness Testimony and Litigation Support, Gasoline Remediation Site and Sewer/House Explosion Case, Winneconne, WI (MACTEC). Doug provided expert witness testimony and investigated the potential causes of and chemical fate and transport mechanisms responsible for a house explosion case. Plaintiffs alleged that vadose and saturated zone petroleum remediation activities at a service station located a few blocks from the residence and subsequent transport of gasoline vapors through a sewer line/backfill were the fuel source for the explosion. He analyzed gasoline vapor transport rates and concentrations in the Douglas J. Cosler, Ph.D., P.E. - Page 3 of 15 subsurface at the service station site, in the 12-inch sewer pipe, and within the sewer backfill. Doug demonstrated that gasoline vapors could not have migrated to the residence between the time that remediation stopped and the house exploded. He also demonstrated that gasoline vapors at explosive levels could not have migrated up the sewer lateral and into the house. His analyses showed that sewer gas (methane) was the likely cause of the explosion because a methane source was present in the sewer line near the residence (sewage blockage due to tree root growth through pipe joints) and lighter -than -air methane naturally migrates upslope along sewer lines and laterals. Remedial Investigation and Feasibility Study (REFS) and Expert Witness Testimony for the Old Southington Landfill Superfund Project, Southington, CT (MACTEC). Doug developed a three- dimensional groundwater flow model (MODFLOW) to evaluate source control alternatives for a municipal landfill that received solid and semi -solid waste materials (primarily VOC). In the vicinity of the landfill, high -permeability deposits in the bottom portion of the aquifer and the presence of a neighboring pond caused large downward groundwater flow components that complicated contaminant transport analysis. He directed the site investigation that focused on the landfill and underlying and downgradient portions of the regional aquifer. He prepared an expert report and provided expert witness testimony for insurance litigation regarding the nature and timing of waste disposal in the landfill. Expert Witness Testimony, Hydrogeologic Investigations of a Gasoline Station, CT (MACTEC). Doug provided expert witness testimony regarding the results of a hydrogeologic investigation to determine the source of petroleum contamination within a telephone company utility conduit. He provided opinions concerning groundwater flow and chemical transport rates in the surrounding aquifer, age dating of petroleum products, and the potential relationship of gasoline -related contaminants in a utility manhole to historical petroleum releases at an upgradient gasoline station. Remedial Investigation, Site Remediation and Expert Witness Testimony, FormerMGP Site, Concord, NH (MACTEC). Historical discharges of carburetted water gas tar contaminated a 10-acre pond and the underlying groundwater with aqueous -phase constituents and NAPL. Contaminants included PAHs and BTEX compounds. Doug designed the hydrogeologic investigation to determine the nature and extent of groundwater and NAPL contamination. He performed data evaluations to assess the potential for vertical and horizontal migration of NAPL and the potential for contamination of a river adjacent to the site. He also prepared two expert reports and provided expert witness testimony for two related insurance litigation actions regarding the timing and ongoing nature of pond contamination and contamination from the former MGP, located upgradient from the pond. Remedial Design Evaluation and Expert Witness Support, Chlorinated Solvent and Petroleum Contamination Site, MO (MACTEC). Doug served as a company expert for litigation involving a groundwater extraction system designed to control LNAPL and aqueous -phase contaminants. Plaintiffs (downgradient property) claimed that the extraction system was not controlling contamination. Doug developed a hydraulic model of the site, analyzed in detail the groundwater capture zone, and demonstrated that the system was very effective in controlling LNAPL and aqueous -phase contaminant migration. Douglas J. Cosler, Ph.D., P.E. - Page 4 of 15 Expert Witness Testimony, Petroleum Contamination Site, Concord, NH (MACTEC). Doug served as a hydrogeology, coal tar, and petroleum fate and transport expert for property damage litigation involving a fuel oil distributor and former MGP site. The plaintiff claimed that coal tar contamination from the former MGP caused environmental damage and increased construction costs for a new hotel being built at the site. Doug performed petroleum transport and fingerprinting analyses and demonstrated that the fuel oil distributor located immediately upgradient from the subject property was the likely source of contamination - not coal tar. Remedial Investigation, Design, and Expert Witness Testimony, Former MGP Site, Laconia, NH (MACTEC). Doug reviewed site investigation reports and evaluated hydrogeologic conditions, contaminant sources, and NAPL mobility at a former MGP site. Historical MGP waste releases (coal tar) had contaminated soil and groundwater and dissolved -phase constituents, and NAPL had migrated into adjacent surface water bodies. He developed conceptual remedial alternatives for the site and evaluated NAPL containment and collection designs. He prepared an expert report and provided expert witness testimony for insurance litigation regarding the timing and ongoing nature of pond contamination. Expert Witness Report, Former MGP Site, Goshen, IN (MACTEC). Provided litigation support and expert report preparation for a case involving a former MGP site. Technical aspects of the project involved hydrogeology, coal tar, and petroleum fate and transport. Remedial Alternatives Evaluation and Expert Witness Report Preparation, Former Electronics Manufacturing Facility, Manchester, NH (MACTEC). Historical releases of tetrachloroethene (PCE) and PCE dissolved in fuel oil caused soil and groundwater contamination at the site. Contaminants were present as dissolved -phase constituents and DNAPL. Doug evaluated data regarding site hydrogeology and contaminant fate and transport to assess the relative contributions of the PCE sources. He evaluated the feasibility and costs of potential remedial alternatives and prepared an expert report assessing the relative contributions of the two different sources of contamination. Groundwater Flow and Aqueous -Phase Chemical Fate and Transport Developed Adaptive Groundwater, a Three -Dimensional Groundwater Flow and Chemical Transport Code based on the Adaptive Mesh Refinement Method (Adaptive Groundwater SolutionsLLC). Adaptive Groundwater is a highly -scalable, three-dimensional numerical code for high -resolution simulation of groundwater flow and solute transport problems. Dynamic adaptive mesh refinement (AMR) and multi- threading are used to automatically generate unstructured grids to handle multiple scales of flow and transport processes. This is done by translating and adding/ removing telescoping levels of progressively finer subgrids during simulation (hiips://www.rockware.com/product/overview.php?id=329). Groundwater Flow, Contaminant Transport, and Biodegradation Model, Feasibility Study and Natural Attenuation Assessment, Estes Landfill Site, Phoenix, AZ (MACTEC). Doug developed three-dimensional groundwater flow and contaminant transport models to simulate current and future, long-term TCE, cis 1,2- DCE, and vinyl chloride (VC) concentrations in the sand and gravel, overburden aquifer at the Estes Landfill site. He used MODFLOW and MT3D99 to simulate chemical transport and fate mechanisms, Douglas J. Cosler, Ph.D., P.E. - Page 5 of 15 including advection, dispersion, dilution by surface water, sorption to soil, and TCE>DCE>VC biotransformation modeled as a sequential, first -order decay -chain process. He computed site biotransformation rates from historical chemical data and transport model calibration. He demonstrated that natural attenuation was a viable remedial alternative, primarily due to significant source -area VOC depletion and high biodegradation rates (reductive dechlorination and direct oxidation of DCE and VC). Combined MTCA REFS and RCRA RFI/CMS Plus Independent Cleanup Actions, Confidential Metals Manufacturing Facility, WA (Hart Crowser). As the Hydrogeologist and Technical Lead for PCB fate and transport issues during work on this large metals manufacturing facility, Doug developed a three- dimensional transport model of the PCB plume that incorporated the variation in mobility and mass fraction of each of the 209 congeners in the PCB mixture. He constructed a three-dimensional groundwater flow/transport model (MODFLOW/MT3D99) to analyze the capture zones and effluent concentration variations for multiple extraction wells with various screened -interval depths. He investigated PCB contamination sources at the site, including industrial wastewater transfer line leaks and unsaturated/saturated zone water contact with contaminated soils. Doug also developed an innovative two-dimensional, rate -limited PCB congener and colloid transport model to evaluate fate and transport mechanisms at the site. The model simulates the transport of all 209 PCB congeners simultaneously, both as aqueous -phase (i.e., dissolved in groundwater) and colloidal (sorbed to mobile colloids flowing with the groundwater) fractions. Colloid filtration due to interactions with the porous media is included. Because of the high groundwater velocities at the site, the model also incorporates rate -limited soil to groundwater chemical partitioning (nonequilibrium chemical sorption) and nonequilibrium groundwater to colloid PCB sorption mechanisms. Remedial Design and Natural Attenuation Modeling, Savage Municipal Water Supply Superfund Site, Milford, NH (MACTEC). Doug developed three-dimensional groundwater flow and solute transport models of this extensive drinking water aquifer using MODFLOW and MT3D. DNAPL releases (PCE and TCA) caused groundwater contamination. Doug directed evaluation of data collected during field permeability testing, monitoring well sampling, and extensive vertical groundwater profiling using microwells. He modeled the effectiveness of various remedial design alternatives that included soil excavation and hydraulic containment in the source area, hydraulic control of downgradient portions of the PCE and TCA plumes, and natural attenuation due to biodegradation, natural groundwater flushing, and dilution by rainwater and river recharge. Doug estimated biodegradation rates using 1) long-term measurements of VOC concentration reductions along the plume centerline, 2) comparisons of parent to daughter compound concentrations, and 3) computations of total VOC mass reductions in the aquifer. In addition, the natural attenuation evaluation used other analytical parameters (e.g., electron acceptor concentrations) to assess the strength of the biodegradation evidence based on the Technical Protocol for Natural Attenuation of Chlorinated Aliphatic Hydrocarbons in Ground Water. He used the MODFLOW model and the AQTESOLV software to analyze the pumping test data. He used AQTESOLV and the Hantush solution for partially -penetrating wells to analyze the single -well tests. Natural Attenuation Software Development, Risk -Based Correction Action (RBCA) Tier 2 Analyzer (MACTEC). Doug authored the commercial software package RBCA Tier 2 Analyzer, a two-dimensional Douglas J. Cosler, Ph.D., P.E. - Page 6 of 15 groundwater flow and biodegradation (transport) model. The software provides five different transport simulation capabilities: 1) single constituent; 2) the PCE>TCE>DCE>VC sequential -decay sequence that occurs during reductive dechlorination; 3) instantaneous BTEX biodegradation with a single electron acceptor (oxygen); 4) instantaneous BTEX biodegradation with multiple electron acceptors (oxygen, nitrate, iron(III), sulfate, carbon dioxide); and 5) kinetics -limited BTEX biodegradation with multiple electron acceptors. The transport model can simulate either equilibrium or non -equilibrium (one-, two-, or multi -site sorption) partitioning between water and soil. The software provides a design tool that can be used for a wide variety of problems, including the analysis of remedial alternatives such as groundwater pump and treat systems (including extraction well concentration "tailing" effects caused by slow contaminant desorption from soil), natural attenuation evaluation, and source remediation level determination. Remedial Investigation and Feasibility Study (RETS) for the Gallups Quarry Superfund Site, Plainfield, CT (MACTEC). Designed investigations of this former waste disposal site to evaluate the nature and extent of groundwater and residual soil (source area) contamination. The initial field program included geophysical surveys, a source -area soil vapor survey, installation and sampling of 50 microwells, wetlands delineation, and surface water/sediment sampling. Doug performed three-dimensional computer visualization of the contaminant plume based on microwell results to direct monitoring well installation. He performed two- dimensional flow modeling to identify an off -site source of groundwater contamination and developed a three-dimensional groundwater flow and chemical transport model (MODFLOW/MT3D) of the site to facilitate the evaluation of remedial alternatives during the FS process. Darling Hill Superfund Site Remedial Investigation and Feasibility Study (RUFS), Lyndonville, VT (MACTEC). As Technical Leader during the Rl/FS for a municipal well field contaminated with VOCs, Doug directed the site investigation, which focused on a disposal area upgradient of the well field and a highly permeable sand and gravel aquifer. The investigation included geophysical investigations, a soil gas survey, boring and well installations, groundwater sampling and analysis, air sampling, surface water and sediment sampling, and pumping and slug tests. Doug developed a three-dimensional analytical groundwater flow model to evaluate potential plume control at the disposal area and the municipal well field. He also constructed a one-dimensional, numerical contaminant transport model, coupled with a chemical leaching model of the waste disposal area, to estimate cleanup times in the regional aquifer in response to various source control alternatives. Evaluation ofNew Monitoring Well Design and Sampling Techniques to Determine Vertical Concentration Variations in an Aquifer, Independent Research Project (MACTEC). Performed independent research to determine new monitoring well designs and sampling techniques that can provide the necessary data to evaluate vertical concentration variations in an aquifer. Doug developed two-dimensional, numerical axisymmetric groundwater flow and chemical transport models to analyze time -dependent monitoring well concentrations during sampling as a function of various vertical concentration distributions in the aquifer and different well designs. The results of this research demonstrated that discrete intervals of monitoring wells with long screens (e.g., 10 to 20 feet or more) can be sampled in a manner that allows both the vertical plume location and concentration variation in the aquifer to be determined. The research also showed that the time vs. concentration responses of a well during a sampling event lasting a few days Douglas J. Cosler, Ph.D., P.E. - Page 7 of 15 exhibit characteristic shapes that can be directly related to aquifer properties and well design parameters and the vertical concentration distribution. He computed a series of concentration vs. time "type curves," analogous to time-drawdown type curves for aquifer permeability tests, that can be matched with measured time -concentration responses. Evaluation and Recommendation of Hydrologic Models for the Department of Natural Resources, Commonwealth ofKentucky (D'Appolonia). Doug performed an extensive analysis of hydraulic/hydrologic simulation models for the Department of Natural Resources, Commonwealth of Kentucky. He evaluated more than 60 hydrologic (i.e., watershed), surface water, and groundwater computer models for simulating flow and contaminant transport that could be used in determining the potential hydrologic and environmental impacts of coal mining operations at various locations in Kentucky. He made several code modifications to the USACE's STORM, Stream Hydraulics Package (SHP), and Water Quality for River/Reservoir Systems (WQRRS) models. Mine Inflow Evaluation for the Shell Minerals Company, IN (D'Appolonia). To evaluate groundwater inflow rates into a 30-square mile underground coal mine in southwestern Indiana during a 30-year mine life, Doug developed a three-dimensional computer model to simulate groundwater flow into various mine panels from an overlying sandstone aquifer by three processes: (1) artesian flow from portions of the aquifer outside of the mine plan area, (2) gravity drainage of water from the voids in the overlying sandstone, and (3) infiltration through a shale layer separating the aquifer and coal seam. Site Investigation and Hydrogeologic Study, Massachusetts Contingency Plan (MCP), Manufacturing Facility (MACTEC). Designed an investigation to characterize the nature and extent of VOC contamination in a shallow overburden -bedrock aquifer system underlying a manufacturing facility. The investigation included soil vapor analysis, overburden and bedrock monitoring well installation, and permeability testing. Doug designed an interim pump and treat system to control contaminant migration from a source area containing PCE in the form of a NAPL. Hydrogeologic Study and Groundwater Remediation for an Industrial Facility, NH (MACTEC). Served as the Technical Leader during the Phase I investigation and performed data evaluation for this 70-acre salvage yard site. The investigation included evaluation of VOC contamination in the groundwater and the design, installation, and operation of a pump and treat system. Doug developed a two-dimensional, axisymmetric groundwater flow model to evaluate the data from a pumping test involving a large -diameter, partially penetrating water supply well. He performed groundwater flow modeling for the final engineering design of the pump and treat system. Design of Waste Disposal Facility for the U.S. Department of Energy, WV (D'Appolonia). Doug designed the leachate collection system for a waste disposal facility that contained process waste from a proposed solvent -refined coal preparation plant near Morgantown, West Virginia. The 800-acre-foot impoundment consisted of two embankments approximately 60 feet in height constructed from coarse refuse, a primary spillway system, a 5-foot clay liner beneath the impoundment, and an underdrain system directly above the liner to reduce the liquid content of the waste and thereby decrease seepage of contaminants through the clay blanket. He performed a detailed computer simulation of the underdrain system performance to Douglas J. Cosler, Ph.D., P.E. - Page 8 of 15 determine the hydrostatic pressure reduction above the clay liner as a function of waste permeability, drain spacing, ground slope, saturated waste depth, and drain dimensions. Remediallnvestigation/Feasibility Study (RI/FS) forAllied Chemical Company, OH(D'Appolonia). Doug performed a groundwater contamination evaluation and remedial design study for a chemical plant bordering the Ohio River. He developed two-dimensional groundwater flow and chemical transport models to evaluate migration beneath a stream to a local municipal well field and computed a groundwater mass balance to determine the percentage of site groundwater flow reaching the well field, the Ohio River, a neighboring stream, and an adjacent property. He used the calibrated model to screen remedial alternatives and determine cleanup levels. Vadose Zone Flow and Transport Vadose Zone and Hydrogeologic Modeling of Storm Water Detention Facilities, Vancouver, WA (Hart Crowser). Doug developed a three-dimensional saturated/unsaturated groundwater flow model of a storm water detention facility using the USGS computer program SUTRA (Saturated -Unsaturated Transport). He dynamically linked the SUTRA code with watershed hydrology (i.e., runoff hydrograph) and detention basin (storage and discharge rate vs. elevation) models. He modified the SUTRA code to incorporate the hydrologic and hydraulic models as subroutines, which provided storm water runoff inflow rates and time - dependent water elevations in the detention basins. Water elevations were converted to time -dependent specified pressure node values in SUTRA. He added transient discharge rates through the porous boundaries of the detention facilities (computed by SUTRA) to the outflow hydrographs. Doug used the models to evaluate the impacts of several factors on the storm water detention facility performance and design, including groundwater table mounding, hydraulic conductivity (K) heterogeneity, the ratio of vertical to horizontal K, detention basin storage capacity, and storm event recurrence interval. Hydrologic Impact Assessment at a Waste Isolation Pilot Plant for the U.S. Department of Energy, NM (D'Appolonia). Evaluated potential salt removal from beneath a radioactive waste disposal facility enclosed in a 2,000-foot-thick salt formation in southeastern New Mexico. The objective was to determine the size and geometry of a dissolution cavity that could form beneath the facility in the next 10,000 years due to hydraulic interaction with a water -bearing unit located 1,000 feet below. Doug evaluated potential mechanisms for salt dissolution and migration to the underlying unit (e.g., diffusion or advection currents produced by density differences), derived analytical equations to quantify the salt removal rate and cavity geometries, and developed a computer model of salt transport in the water -bearing unit. Hydrologic Impact Assessment and Vadose Zone Modeling for the Exxon Minerals Company, WI (D'Appolonia). Evaluated potential hydrologic impacts on the groundwater and surface water regimes due to minerals mining and the related disposal of inorganic wastes at a 400-acre site. Doug developed site -specific computer models of saturated/unsaturated flow and transport to predict changes in groundwater flow rates, water quality, and water levels in hydraulically -connected lakes. He used predictions encompassing an estimated 100-year mine life to negotiate a work plan with the Wisconsin DNR. Douglas I Cosler, Ph.D., P.E. - Page 9 of 15 Remedial Investigation/Feasibility Study, Vadose Zone Modeling, and Remedial Design for a Former Wood -Treating Facility, Olympia, WA (MACTEC). Doug directed the hydrogeologic investigation and remedial design for this wood -treating site. The site involved tidally influenced groundwater contaminated with polynuclear aromatic hydrocarbons (PAH), chlorinated dibenzo-p-dioxins, pentachlorophenol (PCP), LNAPL (PCP carrier oil), and DNAPL (creosote). The investigation consisted of installation of monitoring wells specifically designed to detect LNAPL and DNAPL, aquifer tests, long-term tidal monitoring, salt water intrusion evaluation, aquifer water budget (infiltration) modeling, and treatability studies for bioremediation of soil and groundwater. Doug designed a NAPL and groundwater extraction system and developed a two-dimensional, numerical groundwater flow model as part of the groundwater extraction system design. He performed one-dimensional unsaturated zone vapor transport modeling to estimate leachate and soil gas flux loadings to groundwater. He used the AquiferTest software to analyze the pumping test data and analyzed tidal variations in water -level amplitude and phase lag to evaluate hydraulic conductivity variations. Vapor -Phase Transport Modeling, Lipari Landfill Superfund Site, NJ (MACTEC). Doug constructed a vertical, one-dimensional vapor (soil gas) flux model to calculate VOC emission rates from contaminated soil downgradient of the landfill. He used the emission estimates as source terms in an atmospheric dispersion model to compute air concentrations in the immediate vicinity of the contaminated soil and at several downgradient receptors and used the results to estimate health risks caused by inhalation exposure. These modeling results and health risk estimates provided the necessary data to determine excavation depths for contaminated soil and the thickness of a soil cap that would reduce future exposures to acceptable levels. Health Risk Assessment, Massachusetts Contingency Plan (MCP), Auto Auction Facility (MACTEC). Doug performed the exposure assessment for potential exposure to VOC contamination resulting from a leaking underground fuel tank. He developed a one-dimensional, unsaturated zone, soil gas flux model for estimating indoor air concentrations in domestic buildings overlying subsurface areas contaminated by the spill. He also developed a two-dimensional groundwater transport model for estimating downgradient concentrations beyond the existing monitoring network. Development of Performance Goals for Remedial Measures, a Risk -Based Approach for a Manufacturing Facility, OH (MACTEC). Computed exposure point concentrations for a health risk assessment to determine performance goals for soil and groundwater remediation at a 2-acre site contaminated with several organic chemicals. Doug was responsible for the exposure assessment that involved the development of a groundwater transport model to perform two basic calculations: 1) rate of chemical removal from the contaminated areas of the unsaturated zone soil, and 2) two-dimensional chemical advection and dispersion in the shallow groundwater unit downgradient from the source area. The computation of contaminant removal from the unsaturated zone involved a one-dimensional (vertical) analysis of advection due to infiltration and molecular diffusion through the water and air phases of the soil. He also calculated contaminant dilution in the sand layer using a calibrated two-dimensional (horizontal) transport model. Douglas J. Cosler, Ph.D., P.E. - Page 10 of 15 NAPL Characterization and Modeling Petroleum Extraction System Optimization, Former Manufacturing Facility, MA (MACTEC). Developed and calibrated a two-dimensional fuel oil flow model using the SPILLCAD software to evaluate historical free product recovery volumes and optimize extraction well locations and oil and groundwater pumping rates. He used the calibrated oil flow model to (i) demonstrate the effectiveness of the recovery system in minimizing the future risk of off -site free product transport and (ii) estimate the time period required to obtain the remedial goals for the site. Remedial Design and Investigation, Former Manufactured Gas Plant (MGP) Site, Fort Wayne, IN (MACTEC). Doug designed a groundwater and NAPL (coal tar) containment and collection system for this former MGP site adjacent to a river. A sheet pile wall provided containment along the perimeter of the site, and a trench system with collection pipes and wells collected groundwater and NAPL. Doug developed a three-dimensional groundwater model (MODFLOW) to determine the required water levels in the various collection trench segments to provide hydraulic control of the groundwater plume and optimize NAPL recovery. Remedial Design and Investigation, Former MGP Site, Hammond, IN (MACTEC). Doug designed a NAPL (coal tar) containment system at this former MGP site to prevent NAPL migration into a river that formed the downgradient site boundary. Doug evaluated slurry wall and sheet piling designs. He developed a three-dimensional groundwater model (MODFLOW) of the site to evaluate optimal containment wall designs (e.g., wing wall orientation and length) for minimizing off -site groundwater transport of contaminants. In addition, he used the model to evaluate water level increases on the upgradient side of the wall and potential design options (e.g., gates) to mitigate this effect. Surface Water Modeling Evaporation Prediction for Heated Water Bodies, Research Project for the Electric Power Research Institute, GA (Massachusetts Institute of Technology). Evaluated the evaporative heat loss from a series of heated (70 degrees Celsius) cooling ponds (1 to 5 acres) and canals. Doug developed a one-dimensional hydrothermal model to evaluate the temperature distribution and the energy budgets for the system of water bodies. He performed a literature review of evaporation prediction methods, emphasizing methods capable of predicting combined free (thermally induced) and forced (wind) evaporative heat loss. The research resulted in the formulation of a new evaporation equation that more accurately predicts heat loss from water bodies for conditions, such as high water temperature, where both free and forced evaporation are important. Site Evaluation of Two Nuclear Power Plants for Northeast Utilities, New England (Massachusetts Institute of Technology). Doug evaluated waste heat transport from two nuclear power generation facilities located along the coast of New England. He developed two-dimensional numerical tidal hydrodynamic and Douglas J. Cosler, Ph.D., P.E. - Page 11 of 15 thermal transport models to evaluate temperature increases in adjacent estuaries. He used the temperature simulations to locate new water intakes and to determine heat effects on sediment biota. Sewage Disposal Outfall Siting Study for the Massachusetts Water Resources Authority, Boston, MA (Massachusetts Institute of Technology). Doug evaluated tidal hydrodynamics and contaminant transport in Boston Harbor as part of the design of the new Deer Island sewage treatment plant. He used mass loading data at the existing Deer and Nut Island treatment plants in conjunction with measured concentration distributions for six chlorinated VOCs to calibrate dispersion coefficients and first order surface volatilization rates for the compounds. He used current meter measurements for calibration of the two- dimensional, harmonic hydrodynamic model. He simulated harbor concentrations for several planned diffuser outfall locations using a two-dimensional, transient contaminant transport model that was linked with the hydrodynamic model. Estimate of Toxic Chemical Loadings to Puget Sound, Washington State Department of Ecology Toxics Cleanup Program, WA (Hart Crowser). Doug was the Technical Director assisting Ecology with a multi- year effort to develop strategies, remedial actions, and performance measures to protect and restore the overall health of the Puget Sound ecosystem. He identified toxic chemicals of concern and characterized contaminant sources and pathways (e.g., stormwater runoff, municipal/industrial wastewater effluents, groundwater discharge, chemical spills, and atmospheric deposition). For each of the 17 chemicals of concern, Doug estimated average annual rates of mass loading (runoff rates and stormwater concentrations) to Puget Sound via each pathway. He developed a probabilistic approach to characterizing data uncertainty that involved computing cumulative probability distributions for each mass loading pathway. Surface Water Quality Impact of Treatment System Effluent, In dus tri-Plex Trust, Sup erfund Site, Woburn, MA (MACTEC). Developed two-dimensional hydrodynamic and contaminant transport models of a 10-acre impoundment to evaluate water quality impacts of the treatment system effluent from a series of groundwater extraction wells. Both organic (VOC) and inorganic contaminants were present in the waste stream. Steady-state hydrodynamic simulations, qualitatively verified by field observations, provided an understanding of the velocity distribution in the impoundment that was a function of both tributary and treatment system inflows and large water depth variations of 5 to 20 feet. Doug incorporated depth - averaged contaminant concentration distributions computed using the transport model in an aquatic impact assessment designed to determine preferred effluent discharge locations and rates. Surface Water Quality Impact of Dam Breach, Bangor Hydroelectric, ME (MACTEC). Doug developed a numerical, one-dimensional dissolved oxygen transport model to evaluate receiving water quality impacts from hydrodynamic changes caused by the breaching of a dam in a large river system. He used the USACE's stream hydraulics model HEC-1 to simulate river stage and velocity for a range of breach elevations and stream flow rates. For each flow field the transport model provided estimates of dissolved oxygen changes in the river system. These results demonstrated the beneficial effects of leaving the dam in place. Sedimentation and Erosion Control Plan Design for DuPont, SC (D'Appolonia). Designed the sedimentation and erosion control plan for a 200-acre site disturbed during construction of a waste Douglas J. Cosler, Ph.D., P.E. - Page 12 of 15 processing facility. Evaluations included calculation of storm runoff hydrographs, the design of three sedimentation basins with heights ranging from 10 to 20 feet and storage capacities of 3 to 5 acre-feet, hydraulic design of primary and emergency spillways for the basins, specification of diversion ditch locations and sizes, and design of various other erosion control measures. Research Research Assistant, School of Earth Sciences, The Ohio State University, Columbus, OH, 2003 — 2006. • Dissertation: Numerical Investigation of Field -Scale Convective Mixing Processes in Heterogeneous, Variable -Density Flow Systems Using High -Resolution Adaptive Mesh Refinement Methods. Advisor: Motomu Ibaraki. • Ph.D. Research: Developed new adaptive simulation software for high -resolution, field -scale modeling of non -linear, variable -density ground water flow systems. Examined practical problems such as in situ chemical oxidation of contaminants by dense treatment fluids, water supply applications such as freshwater storage and recovery in coastal aquifers, and saltwater intrusion assessments. The software automatically adjusts to multiple scales of convective mixing processes by translating and adding/removing telescoping levels of progressively finer subgrids to maintain a specified numerical accuracy throughout the global simulation domain. Adaptive mesh refinement methods and higher - order Eulerian-Lagrangian discretization schemes were used to construct a three-dimensional flow and transport code capable of simulating fine -scale (-1-10 cm) instability development and resulting convective mixing in field -scale variable -density ground water flow systems. Because the flow and transport solutions for each subgrid are computed independently, field -scale simulations are broken into multiple smaller problems that can be modeled more efficiently and with finer detail. Convective mixing in heterogeneous porous media is shown to be more amenable to prediction than previously concluded. Convective mixing rates are related to the geostatistical properties of the aquifer (variance and mean of the log permeability distribution, horizontal and vertical correlation scales), the fluid density difference, the magnitude of local small-scale dispersion, the effects of different permeability field realizations, the injection well size and orientation, hydraulic parameters such as injection rate and regional hydraulic gradient, and the spatial resolution. Further, three-dimensional fluid mixing rates are related to mathematical expressions for density -dependent macrodispersivity that are based on stochastic flow and solute transport theory. Colloid transport modeling: Simulated colloid and radionuclide injection experiments for fractured - rock test site in Switzerland. Used two-dimensional finite element code COLFRAC (flow and transport of colloids and contaminants in discretely -fractured porous media) to perform sensitivity analyses involving: fracture aperture, spacing, connectivity; secondary permeability and diffusion rate in rock matrix; equilibrium and kinetic radionuclide sorption parameters for colloids and fracture walls; longitudinal dispersivity; colloid filtration coefficient; and radionuclide decay rate. Research Assistant, Department of Civil & Environmental Engineering, Massachusetts Institute of Technology, Cambridge, MA, 1984 — 1987. • Areas of Specialization: Hydraulics/hydrology, surface water heat transport mechanisms, heat transfer in unstable atmospheric boundary layers, tidally- and density -driven flow/transport, chemical fate/transport, numerical methods (finite element, finite difference, Eulerian-Lagrangian). • Thesis: Evaporation from Heated Water Bodies, Predicting Combined Forced Plus Free Convection. Advisor: Eric Adams. Constructed hydrothermal model to compute evaporative heat loss from 70' C cooling ponds and canals based on simulated temperature distributions and energy budgets. Douglas J. Cosler, Ph.D., P.E. - Page 13 of 15 Formulated new evaporation equation that more accurately predicts heat loss from heated water bodies for conditions where both free and forced evaporation are important. • Surface water modeling: Analyzed tidal hydrodynamics and contaminant transport in Boston Harbor and Massachusetts Bay for design of new Deer Island sewage treatment plant. Constructed two- dimensional, finite element hydrodynamic (harmonic) flow and Eulerian-Lagrangian transport models to evaluate mixing of treatment plant effluent for alternative multi -port diffuser designs and locations. • Hydrothermal modeling: Developed two-dimensional finite element, tidal hydrodynamic and thermal transport models to evaluate waste heat transport in estuaries for two nuclear power generation facilities. Research Assistant, Department of Civil & Environmental Engineering, The Ohio State University, Columbus, OH, 1977 — 1979. • Areas of Specialization: Turbulent transport processes, hydraulics/hydrology, numerical methods. • Thesis: Numerical Simulation of Turbulence in a Wind -Driven, Shallow Water Lake. Advisor: Keith Bedford. Developed three-dimensional hydrodynamics code (finite difference) using large -eddy simulation techniques. Evaluated energy cascade process for turbulent flows in lakes through spectral analysis of velocity fluctuation time series. Independent Research, 1990 — 2002. • Effects of Rate -Limited Mass Transfer, Vertical Concentration Distribution, and Well Design on Ground -Water Sampling and Remediation: Constructed numerical axisymmetric flow and nonequilibrium (multi -rate) transport models to simulate monitoring/extraction well concentrations as a function of plume shape and well design. Showed how sample concentration variations with time can be used to determine vertical concentration distributions in plumes and aquifer properties such as vertical anisotropy ratio, porosity, retardation factor, and soil -water mass transfer parameters. • Commercial Contaminant Transport and Biodegradation Modeling Software: Author of the Risk - Based Correction Action (RBCA) Tier 2 Analyzer, a two-dimensional ground water flow, nonequilibrium (multi -rate) transport, and biodegradation model. Software is based on Eulerian- Lagrangian solution of transport equation with alternating direction implict (ADI) technique for dispersion, and fourth -order Runge-Kutta scheme for PCE decay chain and BTEX biodegradation terms. Teaching Instructor, School of Earth Sciences, The Ohio State University, Columbus, OH, 2006. • Instructor for graduate -level courses in hydrogeology and environmental risk assessment, and undergraduate courses in hydrology and water resources. Teaching Assistant, School of Earth Sciences, The Ohio State University, Columbus, OH, 2003-2006. • Taught several class sessions of graduate -level courses in hydrogeology and environmental risk assessment, and an undergraduate course in water resources. Assisted in the preparation of lecture materials and homework assignments, developed class projects involving field applications, and guided group discussions among students during classes. • Instructor for laboratory sessions of class in earth sciences and water resources. Prepared review materials and lectured on fundamental concepts, and directed students during laboratory exercises. Mathematics Tutor, Boston Partners in Education, Boston, MA, 2001. Douglas J. Cosler, Ph.D., P.E. - Page 14 of 15 • Served as volunteer tutor for high school students in Boston Public School system. Taught individual studies mathematics course in preparation for Massachusetts Comprehensive Assessment System (MCAS) proficiency tests. Teaching Assistant, Department of Civil & Environmental Engineering, Massachusetts Institute of Technology, Cambridge, MA, 1984 — 1987. • Instructed laboratory sessions of undergraduate fluid mechanics course. Conducted laboratory demonstrations and directed students during experiments using various fluid mechanics apparatus. Led field trip to conduct a stream tracer study and evaluate stream hydraulics and dispersion characteristics. Engineering Tutor and Coordinator, College of Engineering, The Ohio State University, 1974 — 1977. • Tutored undergraduate engineering students in mathematics, physics, chemistry, and engineering mechanics. Served as student program coordinator responsible for evaluating undergraduate educational requirements, and tutor assignments and schedules. PUBLICATIONS Cosler, D.J. 2006. Numerical Investigation of Field -Scale Convective Mixing Processes in Heterogeneous, Variable -Density Flow Systems Using High -Resolution Adaptive Mesh Refinement Methods. Ph.D. Dissertation, The Ohio State University, School of Earth Sciences, Columbus, Ohio. Cosler, D.J. 2004*. Effects of Rate -Limited Mass Transfer on Water Sampling with Partially Penetrating Wells. Ground Water 42, no. 2: 203-222. Cosler, D.J. 2000. Risk -Based Correction Action (RBCA) Tier 2 Analyzer, Two -Dimensional Groundwater Flow and Biodegradation Model, Ref. Manual. Waterloo Hydrogeologic, Inc., Waterloo, Ontario, Canada. Cosler, D.J. 1997*. Ground -Water Sampling and Time -Series Evaluation Techniques to Determine Vertical Concentration Distributions. Ground Water 35, no. 5: 825-841. Adams, E.E. and Cosler, D.J. 1990*. Evaporation from Heated Water Bodies: Predicting Combined Forced Plus Free Convection. Water Resources Research 26, no. 3: 425-435. Adams, E., Kossik, R., Cosler, D., MacFarlane, J., and Gschwend, P. 1990. Calibration of a Transport Model Using Halocarbons. Estuarine and Coastal Modeling, M.L. Spaulding, ed., ASCE, New York, N.Y., pp. 380-389. Andrews, D.E. and Cosler, D.J. 1989*. Preventing and Coping with Water Pollution. Journal of Testing and Evaluation, ASTM 17, no. 2: 95-105. Walton, R., Kossik, R., Adams, E., and Cosler, D. 1989. Far -Field Numerical Model Studies for Boston's New Secondary Treatment Plant Outfall Siting. Third National Conference on Hydraulic Engineering, New Orleans, Louisiana, August 14-18. Adams, E.E. and Cosler, D.J. 1988*. Density Exchange Flow Through a Slotted Curtain. Journal of Hydraulic Research 26, no. 3: 261-273. Adams, E.E. and Cosler, D.J. 1987. Predicting Circulation and Dispersion Near Coastal Power Plants: Applications Using Models TEA and ELA. Massachusetts Institute of Technology Energy Laboratory Report No. MIT -EL 87-008, 113. Adams, E.E., Cosler, D.J., and Helfrich, K.R. 1987. Evaporation from Heated Water Bodies: Analysis of Data from the East Mesa and Savannah River Sites. Civil Engineer Degree Thesis, Massachusetts Institute of Technology, Cambridge, Massachusetts. Cosler, D.J. and Snow, R.E. 1984*. Leachate Collection System Performance Analysis. Journal of Geotechnical Engineering, ASCE 110, no. 8: 1025-1041. Snow, R.E. and Cosler, D.J. 1983. Computer Simulation of Ground Water Inflow to an Underground Mine. In Proceedings of the First Conference on Use of Computers in the Coal Industry, AIMS, (Y.J. Wang and R.L. Sanford editors), pp. 587-593. W. Virginia University, August 1-3. Douglas J. Cosler, Ph.D., P.E. - Page 15 of 15 Cosler, D.J. 1979. Numerical Simulation of Turbulence in a Wind -Driven, Shallow Water Lake. M.S. Thesis, The Ohio State University, Columbus, Ohio. * Denotes peer -reviewed journal. PRESENTATIONS Cosler, D.J. 2015. An Intelligent Graphical User Interface for MODFLOW and MT31) based on Dynamic Adaptive Mesh Refinement Methods. MODFLOW and More 2015 Conference. Colorado School of Mines, Golden, Colorado, May 31 - June 3. Cosler, D.J. 2013. Numerical Simulation of Multiscale Transport Processes in Variable -Density Flow Systems Using High -Resolution Adaptive Mesh Refinement Methods. MODFLOW and More 2013 Conference. Colorado School of Mines, Golden, Colorado, June 2-5. Cosler, D.J. 2006. Numerical Investigation of Field -Scale Convective Mixing Processes in Heterogeneous, Variable -Density Flow Systems Using High -Resolution Adaptive Mesh Refinement Methods. Geological Society of America Annual Meeting, October 22-25, Philadelphia, Pennsylvania. Cosler, D.J. and Ibaraki, M. 2006. Numerical Investigation of Multiple-, Interacting -Scale Variable - Density Ground Water Flow Systems. American Geophysical Union, Western Pacific Geophysics Meeting, July 24-27, Beijing, China. Cosler, D.J. and Ibaraki, M. 2005. Numerical Investigation of Multiple-, Interacting -Scale Variable - Density Ground Water Flow Systems. Geological Society of America Annual Meeting, October 16-19, Salt Lake City, Utah. Cosler, D.J. 2003. Modeling the Effects of Multirate Mass Transfer on Water Sampling with Partially - Penetrating Wells. Geological Society of America Annual Meeting, November 2-5, Seattle, Washington. Duke Energy Memorandum Regarding CAMA Requirements Introduction The purpose of this document is (1) to establish Duke Energy's compliance with the groundwater assessment and corrective action requirements of the North Carolina Coal Ash Management Act ("CAMA"), and (2) to identify information relevant to the Department's assessment and prioritization of coal ash surface impoundments for closure under CAK8A. Asexplained further below, Duke Energy has submitted all groundwater information required by CAMA to date and will continue to submit information required bythe Department pursuant ioCAK0Aauthority. /\maresult, there ionnbasis for any finding by the Department that Duke Energy has failed tocomply with CAMA. Further, the information submitted by Duke Energy, supplemented by other available, relevant information, is sufficient for the Department to make an evidence -based assessment of the factors that CAMA requires for impoundment prioritization; as a result, it would be legal error for the Department to prioritize the surface impoundments without full consideration of, and findings offact on, each ofthe factors. U. Compliance with Groundwater Assessment and Corrective Action CAMA's groundwater assessment and corrective action provisions are located at North Carolina General Statutes § 130A-309.211. Duke Energy has complied with subsections (a) and (b)aefollows: A. Suboection(o)-GmnundvvoterAmomaament Subsection (a) requires Duke Energy to, at each of its surface impoundments, do three things.. (1) submit a proposed Groundwater Assessment Plan for approval by the Department, (2) begin implementing a Groundwater Assessment Plan approved by the Department, and (3) submit a Groundwater Assessment Report describing all exceedances of groundwater quality standards associated with the impoundment. Duke Energy has moteach ofthese requirements. As you are aware, Duke Energy submitted draft Groundwater Assessment Plans for all of its surface impoundments inNorth Carolina onSeptember 26.2014. The Department provided comments on November 5, 2014, and Duke Energy submitted revised Groundwater Assessment Plans on December 30.2D14. The Department conditionally approved the Plans on various dates earlier this year, NCOENRO194488 and Duke Energy began implementing each plan within 1Ddays ufapproval. Groundwater Assessment Reports describing all exceedances of groundwater quality standards associated with the various surface impoundments were submitted tothe Department inAugust and September. The Department's approval of the plans reflected a determination that the plans met CAMA requirements. Duke Energy's implementation of the plans, including the conditions of approval, under the Department's close oversight, further supports a conclusion that the requirements of Subsection (a) were met. Aoonrdingto8)ep|ain|anguagenf8ubaaodon(a).Ouk*EnerAy'snom[dkannewbh8)o requirements does not depend on the substantive content of the Groundwater Assessment Reports. Duke Energy was required to make plans to assess various groundwater factors, which it did. The Department approved the plans, thereby determining that the Plans would assess those groundwater factors. Duke Energy diligently implemented the Plans. There isnofurther requirement inSubsection (a), or anywhere else in CAMA, that the results of the groundwater assessments definitively establish or disprove the existence ofany condition otasite. |nfact, CAK8Aanticipates that groundwater assessments performed under Subsection (a) may not supply all the information desired by the Department —Subsection (b)(a) requires Duke Energy to include in a proposed Groundwater Corrective Action Plan ^[a]nyother infonnationne|atodtogroundvvaderauooaoman{naquirodbythaDepartmont.^ Had the General Assembly anticipated that Groundwater Assessment Reports would be definitive documents, there would have been no need to authorize the Department to request additional information in the proposed Corrective Action Plans. B. Submeudon(b)-CorrectivaAction Similarly, Subsection (b)requires Duke Energy Lodntwo things: (1)submit aproposed Groundwater Corrective Action Plan, and (2) begin implementing the Groundwater Corrective Action Plan once i1has been approved bythe Department. The deadline for completion ofthe first requirement has not yet passed. The Department and Duke Energy agreed that the Corrective action plans would be submitted in two parts, and Duke Energy has submitted the first part for all fourteen sites with surface impoundments The deadline for submission ofthe second part has not yet arrived. NCOENRO194489 The Corrective Action Plans contain each of the elements from Subsection (b)(1) that were to be included inthe first part submittals. The Corrective Action Plans were prepared byqualified professionals and contain work performed tothe industry standard. Additional information will besubmitted inthe part two submittals. It is premature to evaluate Duke Energy's compliance with this requirement until the submittals are complete. Ui Prioritization mfSurface Impoundments Under CAMA, the Department is charged with developing proposed classifications of surface impoundments according tothe procedures in Worth Carolina General Gtututea§ 13OA,300.213. The prioritization must be based on "a site's risks to public health, safety, and welfare; the environment; and natural resnuroes.^ N.C.Gen. Gtat. §13OA-30Q213(a). |nassessing the risks, the Department must evaluate groundwater data submitted under § 13OA-309.21 1, discharge information submitted under § 130A,38Q.212.and any other information deemed relevant. Further, the Department must consider all of the following: no Any hazards topublic health, safety, orwelfare resulting from the impoundment. "o The structural condition and hazard potential nfthe impoundment. uo The proximity of surface waters to the impoundment and whether any surface waters are contaminated or threatened by contarnination as a result of the impoundment. co Information concerning 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 and all significant factors affecting contaminant transport. on The location and nature of all receptors and significant exposure pathways. oo The geological and hydroOnn|oOira|features influencing the movement and chemical and physical character ofthe contaminants. oo The amount and characteristics of coal combustion residuals in the impoundment. on Whether the impoundment is located within an area subject to a 1 00-year flood. 'o Any other factor the Department deems relevant tuestablishment ofrisk. The Department must issue written declarations, including findings uffact, documenting proposed risk classification. NCOENRO194490 This section requires the Department to make decisions based on the available evidence regarding each ofthe listed factors.' hwould not boconsistent with this requirement for the Department to make a classification decision based solely on one factor and disregard valid information about the others. Further, the section anticipates that the Department will make decisions before complete information about a site is available. For example, it does not require the Department to oump|o{o|y know the vertical and horizontal extent of soil and groundwater contamination for each site; rather it requires that the Department consider infort'nation conceming the vertical and horizontal extent. Similarly, i1does not require the Department to know all factors that might conceivably affect contaminant transport orall conceivable exposure pathways, it requires the consideration only of significant factors affecting contaminant transport and significant exposure pathways. Additional support for this conclusion is found in the fact that this section defines an iterative process by which evolving data, review, and commentary are used to classify surface impoundments as |ow, intennediate, or high risk. This iterative process begins with e provisionally proposed classification by the Department by December 31, 2015 and extends for a minimum (no maximum) of six months while feedback and additional data are received and evaluated by the Department and the Coal Ash Management Commission. Taken as a whole, this section requires the Department to make evidence -based decisions using the best available information inthe record. Duke Energy has submitted substantial evidence into the administrative record. Any classification should be based on this evidence, with the understanding that additional information requests may be relevant to the degree of certainty in the classification but do not undermine the validity ofthe classification. IV. Conclusion The Groundwater Assessment Plans, Groundwater Assessment Reports, and Groundwater Corrective Action Plans submitted by Duke Energy to the Department meet the requirements of CAMA and provide vast data, analysis, and findings. Chief among the findings inadetermination bylicensed 1 Aside from CAMA, the North Carolina Administrative Procedure Act requires that agency decisions be supported "by substantial evidence ... in view of the entire record as submitted." G.S. 1 50B-51 (b)(5). NCOENRO194491 environmental geologists that none of the sites pose an imminent hazard to human health or the environment. Duke Energy is committed to meeting the Department's expectations by providing additional data, fully leveraging the time provided by CAMA's iterative process to ensure final classifications reflect the best science and engineering. Nonetheless, Duke Energy has complied with all ofCAMA'srequirements to date, and available information is sufficient for the Department to develop classifications as required by CAK4A. NCOENRO194492 3 -12-15 NCDENR Conditional Approval of Revised Groundwater Assessment Work Plan AFFFA 74i'AT NR North Carolina Department of Environment and Naturai Resources Pat McCrory Governor March 12, 2015 Mr. Harry Sideris Senior Vice -President Environment, Health, and Safety Duke Energy 526 South Church Street Mail Code EC3XP Charlotte, NC 28202 Re: Marshall Steam Station NPDES Permit No. NCO0O4987 — Catawba County, North Carolina Conditional Approval of Revised Groundwater Assessment Work Plan Dear Mr. Sideris: Donald R. van der Vaart Secretary On December 31, 2014, the Division of Water Resources (Division) received the revised. Groundwater Assessment Plan (GAP) for the above listed facility. The revised GAP was submitted in response to the DWR's Review of Groundwater Assessment Work Plan letter dated November 4, 2014. A review of the plan has been completed and several deficiencies or items requiring clarification were noted. Therefore, in order to keep the site assessment activities on a timely schedule, the Division has approved the revised GAP under the condition that the following deficient items are addressed in the Groundwater Assessment Report; • Comment Section 5.3 Hydrogeologic Site Characteristics: The initial site conceptual model (ISCM) section of the revised GAP does not provide a clear, cohesive description of how constituents of potential concern (CQPCs) may migrate from the source(s) to the receptors through various pathways. It is acknowledged that there is information available to develop an ISCM, but the data are not presented in a manner such as groundwater elevation maps, geologic maps, cross -sections that depict detailed site conditions, flow diagrams, or in a tabulated format to illustrate where data gaps may exist. Duke Energy should incorporate all existing data at the site and be prepared to collect additional data if the Division. determines that additional data gaps exist. Continued site conceptual model development should follow guidelines similar to those presented in the American Standards Testing Measures E1689 - 95(2014) Standard Guide for Developing Conceptual Site Models for Contaminated Sites to direct data collection, data interpretation, and model development efforts. 1611 Maid Service Center, Raleigh, North Carolina 27699-1611 Phone: 919 707-90001 Internet: http:Nwww.ncwater,orgi An Equal Opportunity' Affrrnatwc Action Employer — Made in part by recycled paper NCDENROO68627 Marshall Steam Station March 12, 2015 Page 2 of 3 a Comment 7.1.4 Bedrock Monitoring Wells: The Division suggests installing bedrock monitoring wells at several locations in order to provide data for assessment of multiple flowpath transects across the site, A bedrock monitoring well is recommended near the compliance boundary in the vicinity of well cluster MW-14S/D based on analytical results obtained from those locations that suggest a possible coal ash signature. Installation of a bedrock monitoring well in the vicinity of the MW-7S/D, MW-8S/D, and MW-9S/D well clusters is recommended in order to provide a monitoring location downgradient of the active ash basin adjacent to Lake Norman, In addition, bedrock monitoring wells are recommended in the immediate vicinity of proposed well clusters at AB-6S/D, AB-15S/SL/D, and AL-2S/D so bedrock groundwater can be characterized in the ash basins and landfill. 0 Con-iment Section 7.2 Groundwater Sampling and Analysis: Direction provided in the EPA Region I Low Stress Purging and Sampling Procedure for the Collection of Groundwater Samples from Monitoring Wells (2010) should be followed strictly and any deviations from the procedure must be approved by the Division and documented accordingly. For example, samples should not be collected until pH is stabilized within ± 0,1 for three consecutive readings rather than ± 0.2 written in the GAP. Temperature and specific conductivity readings should stabilize within 3% for three consecutive readings be -fore samples are collected instead of 10% noted in the GAP, Also note that if the pumping rate is so low that the flow-through-cell/chamber volume Qannot be replaced in a 5 minute interval, the time between measurements should be increased accordingly. Speciation of inorganics; from groundwater samples should be focused on wells strategically located along flowpath transects. Collection of data along multiple flowpaths will refine the assessment of water geochemistry and development of flow and. transpo.1.1 models. The Division recommends analyzing samples for radionuclides from wells monitoring well locations MW-14S/D because of elevated concentrations of boron, iron, manganese and sulfate detected during previous sampling events. Other well locations recommended for collecting and analyzing samples for radionuclides include MW-6S and MW-7S so groundwater can be assessed for these constituents beneath the landfill and doNAMgradient of the active ash basin, respectively. Comment Table 5 — Groundwater, Surface Water, and Seep Parameters and Constituent Analytical Methods: Low level Vanadium listed as having a detection limit unit of mg/L. This is likely just a typographical error but the units should be in gg/L rather than mg/.L. Comment Section 7.3.1 Surface Water Samples, 73.2 Seep Samples and 7.2.2 Speciation of Select Inorganics: The GAP text indicates that review of the NCDENR. March 20:14 seep and surface water sampling analytical data will be incorporated into assessment plans to evaluate seep and surface water sample locations at the facility, NCDENR0068628 Marshall Steam Station March 12,2015 Page 3 of 3 s Comment 7.9.2 Development of Kd Terms It is expected that additional solid phase samples will be collected and analyzed for lid determination as well as physical pro -per -ties at strategic locations along flowpath transects. These data will refine the assessment of water geochemistry and development of flow and transport models. Locations where NCDENR's March 2014 seep and surface water sampling data indicated exceedances or elevated concentrations of iron, manganese and other constituents of concern should be incorporated into the assessment's seep/surface water sampling plans with speciation of analytical data sufficient to support delineation and modeling efforts. In addition, technical direction that will serve as the basis of expectations for completion of the site assessment is provided at Attachment 1. Failure to address the deficient items stated above will result in Duke Energy not being in compliance with the stated statutes. Per G.S. 130A-309.209(a) (3) and (4), you must begin implementation of the revised GNP on March 22, 2015 and the Groundwater Assessment Report is due on September 8, 2015, It is our understanding that Duke Energy may have to obtain additional permits to facilitate installation of certain monitoring wells. In the event permits are needed for this purpose, Duke Energy should take all steps necessary consistent with the law to avoid delaying completion of the assessment report. If you have any questions, please contact Bruce Paiiis at (704) 2' ) 5 -2185, Sincerely, 're"Y 4ZeZman, P.G., Director, Division of Water Resources cc: WQROS — MRO WQROS Central Files DENR Secretary - Don van der Vaart HDR (Attn: William Miller) 440 South Church Street, Suite 1000, Charlotte, NC 28202 NCDENR0068629 Attacrtment 1 Page 1 of 6 The items identified in this Groundwater Assessment Plan (GAP) review summary are provided for general discussion for the various parties to agree upon technical direction and content in the revised GAps, comprehensive site assessments 4CSAsj, and corrective action plans (OAPs). is ; � _ '•. a. 1.. A schedule for continued groundwater monitoring is mandated by the Coal Ash Management Act 2014. An interim plan should include at least two rounds of groundwater samples collected' and analyzed in 2015. The analytical results of the first round of data collected in 2015 would be included in the CSA report, while the results of the second round would be submittedas a CSA. addendum. After CSA data can be evaluated, a plan for continued groundwater monitoring can. be developed for implementation in 2016. 2. Sites impacted by inorganics are typically managed using a tiered site analysis which includes four elements as referenced in EPAJ600/R-07/139: Demonstration of active contaminant removal from groundwater & dissolved prime stability; * Determination of the mechanism and rate of attenuation; 8 Determination of the long-term capacity for attenuation and stability of immobilized contaminants, before, during, and after any proposed remedial activities; and * Design of performance monitoring program, including defining triggers for assessing the remedial action strategy failure, and establishing a contingency pl'an,. This reference and the framework described above should' be used as applicable to meet the corrective action requirements found' in 1 aA NCAC 02L .0106.. 3.. Because of uncertainty concerning the site's ability to attenuate contaminants over the long terra given potentially changing geochernical conditions, there is a need to address the elements of the tiered site analysis described above and collect appropriate samples as part of the CSA, GAP development, and continued groundwater monitoring. 4. The Division of Water Resources (Division) Director is responsible for establishing background levels for CQPCs in groundwater. This determination is based on information and data provided by the responsible party and may include formal statistical testing using background' wells with at least four rounds of data. Wells identified as "background" are subject to periodic review based on a refined understanding of site chemistry and hydrogeologic conditions. In general, each facility roust have a background well or wells screened or open to each of the dominant flow systems that occur at the site and are associated with groundwater contamination. Any questions concerning adequacy of background monitoring Locations or conditions at the facilities should be directed to the Regional offices. NCDENR0068630 Attachment I Page 2 of 6 5. Delineation of the groundwater contaminant plume associated with coal combustion residuals is a requirement of the investigation and if off -site monitoring wells are ultimately requiredto perform this task, then it is expected that these activities will be completed as part of the groundwater assessment activities and included in the final report. Documentation of the effort to gain off -site access, or right of way permits, will be provided If off -site access is denied or alternate means of assessing the area were not available within the allocated' timeframe (such as within right-of-ways). Data Requirements and Sampling Strategy 1. Robust data collection is warranted to support timely completion of site assessments and subsequent corrective action plans because of the impending deadlines for completion of CSAs and CAPS, scale and geologic complexity of the sites, the challenges of modeling heterogeneous systems, and .site proximity to potential human and sensitive ecosystem receptors. 2. Robust data collection will be focused along strategically positioned flowpath transect(s) a from ash pond source to potential receptor --as an efficient approach for model development (analytical, geochemical, groundwater flow, and transport) in support of risk assessment and CAP development. Data collected to support evaluation of site conditions along the flowpath. transacts should be located along or defensibly proximate to the modeled transects. I The dataset developed along proposed flowpath transects will include any information needed to determine constituent concentrations, conduct Kd tests, and perform batch geochemical modeling in multiple flow horizons as appropriate. This data will include a) solid phase sample collection for Kd measurement and batch geochemical modeling, inorganic analysis and speciation, and other parameters identified in General Comment #4 of the November 4, 2014 GAP comments issued by DWR, b) solution phase sample collection for total and dissolved inorganic analysis of total concentrations, small pore filtration for dissolved samples, etc,, and c) slug, constant/falling head, and packer testing. The solid phase sample mineralogy, total concentration results, re-dox measurements, and other geochemical parameters will be used as input for equilibrium speciation calculations of redox sensitive constituents calculated by PHREEQC or similar program (EPA/540/5-92%018). This geochemical modeling will: be performed' to identify potential mineral phases, estimated species speclation and concentrations, and will be performed varying key solubility controlling parameters to predict mineral phases, speciation, and concentrations under varying conditions. Solid samples for Kd tests collected from along from proposed flowpath transacts will be handled and preserved in order to eliminate exposure to ambient air in the field. Kd samples should' be collected in plastic bags and sealed with a conventional vacuum plastic bag sealer. The samples will be then placed on ice in a cooler for transport and kept out of direct sunlight. Once received by the analytical laboratory, the Oxidation -Reduction Potential (ORP) of the sample will be measured using, an ORP probe and meter in accordance with ASTM method' G200-19 (Reapproved 2014), Rased on this ORP measurement, either normal or "glove -box" processing of a sample will be applied NCDENR0068631 Attachment 1 Page 3 of 6 (EPA1600/R-06/112). An additional sample will be retained, pending confirmation of subsequent ©RP and DO testing. ORP and dissolved oxygen will be measured in the groundwater monitoring wells subsequently installed at these sample locations. In the event that the groundwater field - measured ORP and DO reveal reducing conditions, the additionally -retained: sample will be subject to glove box processing for the Kd analyses. Refer to EPA/600/R-07/139 Section III for the data collection and characterization needed to support the four -tiered analysis discussed above. 4. Speciations for groundwater and surface water samples should include Fe, Mn, and any COPCs whose speclation state may affect toxicity or mobility (e.g. As, Cr, Se, or others if applicable), This speciation will apply for groundwater samples collected at wells located along proposed flowpath transacts and in wells where these constituents exceed 2L groundwater standards as well as for surface water samples collected within ash impoundments, 5, Solid phase samples shall be analyzed for: minerals present, chemical composition of oxides, hydrous Fe, Mn, and AL oxides content; moisture content; particle size analysis; plasticity; specific gravity; porosity; permeability, or other physical properties or analyses needed to provide input to a chosen model. These analyses for physical properties will be conductedi at up to 15 locations along proposed flowpath transacts where Kd samples are collected, Solid phase samples at up to 15 additional locations will be collected and analyzed for hydrous ferric oxide (HFO) content. At these additional locations where HFO content is analyzed, analyses for physical properties will not be performed. Solid phase samples will be analyzed for total organic content from the same locations where samples are collected for Kd determination. Solid phase samples will be analyzed for total organic carbonate content from the same locations where samples are collected for Kd determination only at facilities located in the coastal plain. 6. In addition to conducting the SPILP leachable inorganic compounds analysis for selected ash samples to evaluate the potential for leaching of constituents to groundwater, the leachable analysis should also be conducted for some soil samples from locations beneath the ash ponds, within the plume, and outside the plume to evaluate potential' contributions from: native soils. 7. In addition to collecting solid phase samples onsite for Kd procedures, soil samples should be also collected from unaffected soils within groundwater flow pathways to evaluate Kd(s) or hydrous ferrous oxide. 8. Rock samples for laboratory analyses should be collected as commented: in General Comment 4 of the November 4, 2014 GAP comments issued by DWR. This GAP review comment indicated that the sample(s) collected from bedrock well soil and rock cores shall be analyzed, at a minimum, for the following: type of material, formation from which it came, minerals present, chemical composition as oxides, hydrous Fe, Mn, and Al oxides content, surface area, moisture content, etc,; however, these analyses were not mentioned in the GAP. The Division: reserves the right to request analysis for organic carbon content, organic carbonate content, and ion exchange capacity if needed to complete the site assessment process, 9. The coal ash and soil analyte lists should match the groundwater analyte lists. 10, Total uranium analysis should be analyzed where total radium is analyzed for groundwater, NCDENR0068632 Attachment I Page 4 of 6 IL If analytical results from a seep sample exceed 2L standards, then the area in: the vicinity of the sample location should be investigated for groundwater contamination. If analytical results from a surface water sample exceed 2B standards, then the area in the vicinity of the sample location should be investigated for groundwater contamination. 12. Surface water/seep samples should be collected during baseflow conditions and that the groundwater monitoring (water levels and sampling) should occur at about the same time. 13. Measurement of strearnflow in selected perennial streams is expected as needed in support of simulation/calibration of flow and transport models; major rivers that serve as groundwater divides are not included in this expectation. Conceptual Model Elements In the CSA report, data gaps remaining should be specifically identified and: summarized. Site heterogeneities should be identified and described with respect to: a) their nature, b) their scale and density, c) the extent to which the data collection successfully characterizes them, d) how the modeling accounts for them, e) and how they affect modeling uncertainty, The impact of data gaps and site heterogeneities should be described in relation to the elements developed in the Site Hydrogeologic Conceptual Model and Fate and Transport Model subsections. 4. For sites in the Piedmont or Mountains, the CSA Report should include a subsection within the Site Geology and Hydrogeology Section titled 'Structural Geology'. This section should describe, a) foliations, b) shear zones, c) fracture trace analysis, and d) other structural components anticipated to be relevant to flow and contaminant transport at the site, S. Duke Energy will include a poster -sized sheet(s) (ANSI E) combining tabulated analytical assessment results (groundwater, surface water, and leachate samples); multiple sheets may be needed to present the data, This should be provided in addition to the Individual analytical results tables that will be prepared for the CSA reports, Any questions concerning format or content of the analytical result summaries should be directed to the Regional offices. Geochemical Modeling 1. The Division agrees that a geochemical model "coupled" to a 3-D fate and transport model is inappropriate given the size and complexity of the sites and the extremely large amount of data required to calibrate such a model. Rather, a 'batch" geochemical model approach should be sufficient for successfully completing the site assessment and/or corrective action plan, 2. Samples collected for 'batch" geochemical analysis should be focused along or defensibly proximate to flowpath transects. 3. To support successful batch geochemical modeling, dissolved groundwater samples collected along a contaminant Towpath transect should be obtained using a 0.1 um filter. This will help ensure a true dissolved phase sample. Note that the dissolved samples are for assessment purposes only and may not be used for purposes of compliance monitoring. If there is uncertainty about which areas/wells will be used in the batch geochemical modeling, the initial round of assessment sampling at the facility can utilize the 0.45 urn filter until the contaminant NCDENR0068633 Attachment 1 Page 5 of 6 flow path transects are selected. Once determined, Duke Energy can go back and re -sample the wells needed for geochemical modeling using the 0.1 um filter. It is recognized that the use of a 0.1 um filter will be difficult for wells with elevated turbidity; in this case, it is recommended that Duke Energy use two filters in series (the water initially passes through a 0.45 um filter to remove larger particles prior to passing through the 0.1 um filter). information for a disposable 0.1um field filter designed specifically for sampling groundwater for metal analysis is provided at the following link: http://www.vosstech.com/index.php/products/filters. If field comparisons of 0.1 versus 0.45 micron filtration at several transect wells at a given site show no significant differences between the two methods, then 0.45 micron filters may be used for evaluating the dissolved phase concentrations at that site. 4. In support of the objectives of General Comment #2 of the November 4, 2014 GAP comments issued by DWR, Duke Energy should add a column titled `relative redox'to the analytical results tables to record the geochemical conditions for that location/sample date. The redox determination should be based on observed DO, ORP, and any other relevant measures and presented for historic and new samples (wells, ash pore water, surface waters, etc.). Relative redox designations may include "iron reducing", "sulfate reducing", mildly oxidizing, moderately oxidizing, etc. and should be footnoted with a statement about the degree of confidence in the designation based on amount and quality of available data. 5. Duke Energy shall also evaluate: a) spatial geochemical trends across the facility and along selected flow paths, b) temporal geochemical trends where observable (such as for compliance boundary wells), along with the likely reason for the change (e.g. increase in seasonal recharge, pond de -watering and subsequent reversal of groundwater flow direction, inundation of well from river at flood stage, etc.) in support of the CAP. This evaluation step will require a comparison of geochemical conditions over time with rainfall data, notable ash capping, dewatering, disposal/removal, or other plant operations, etc. The quality of existing geochemical data will be evaluated using field notes, calibration records, and consistency in redox measurements (e.g. eH vs. raw ORP). Groundwater Models 1. The technical direction for developing the fate and transport modeling will follow guidelines found in Groundwater Modeling Policy, NCDENR DWQ, May 31, 2007, and discussions conducted between Duke Energy and their consultants with the Division. Ultimate direction for completion of fate and transport models will be provided by the Division. 2. The CAP Report should include a subsection within Groundwater Modeling Results titled `Site Conceptual Model' that succinctly summarizes, for purposes of model construction, the understanding of the physical and chemical setting of the site and shall include, at a minimum: a) the site setting (hydrogeology, dominant flow zones, heterogeneities, areas of pronounced vertical head gradients, areas of recharge and discharge, spatial distribution of geochemical conditions across the site, and other factors as appropriate), b) source areas and estimated mass loading history, c) receptors, d) chemical behavior of COPCs, and f) likely NCDENR0068634 Attachment I Pace 6 of 6 Z3 retention mechanisms for COPCs and how the mechanisms are expected to respond to changes in geochemical conditions. 3. Modeling will be included in the Corrective Action Plan (CAP). The four -tiered analysis previously referenced and appropriate modeling should be conducted, and the mass flux calculations described in the EPA/600/R-07/139 should be performed. 4. The CAP Report shall provide separate subsections for reporting groundwater flow models and fate and transport models. 5. The CAP Report should include subsections within Groundwater Modeling Results titled 'Groundwater Model Development' that describes, for each chosen model: a:) purpose of model, built-in assumptions, model extent, grid, layers, boundary conditions, initial conditions, and others as listed in Division guidance. Include in this section a discussion of heterogeneities and how the model(s) account for this (e.g. dual porosity modeling, equivalent porous media approach, etc.). Separate subsections should be developed for the groundwater flow model, fate and transport model, and batch geochemical models, respectively, 6. CAP Reports should include a subsection within: Groundwater Modeling Results titled 'Groundwater Model Calibration' that describes, for each model used, the process used to calibrate the model, the zones of input and calibration variables (for example, hydraulic conductivities) that were used, the actual (measured) versus modeled results for all key variables, and others, Separate subsections should be developed for the groundwater flow model, fate and transport model, and batch geochernical model(s), respectively. 7. CAP Reports should include a subsection within Groundwater Modeling Results titled 'Groundwater Model Sensitivity Analysis' that describes, for each model used, the process used to evaluate model uncertainty, variable ranges tested, and the key sensitivities. Separate subsections should be developed for the groundwater flow model, fate and transport model, and batch geochemical model(s), respectively. Development of Kd Terms 1. Kd testing and modeling in support of CAP development should: include all COPCS found above the NCAC 15A 02L.0106(g) standards in ash leachate, ash pore water, or compliance boundary well groundwater samples. The selected Kd used in transport modeling often will profoundly affect the results. Duke Energy should acknowledge this concept and document within the transport modeling section(s) of the CAP all widely recognized limitations inherent in the estimation of the Kd term. Provide references for guidance and potential sampling methodology related to conducting a baseline ecological risk assessment or habitat assessment, if warranted. NCDENR0068635 10-2-15 NCDENR Draft Comprehensive Site Assessment Comments Pat McCrory Governor October 2, 2015 Mr. Harry Sideris Senior Vice -President Environment, Health, and Safety Duke Energy 526 South Church Street Mail Code EC3XP Charlotte, NC 28202 Re: Marshall Steam Station NPDES Permit No. NC0004987 — Catawba County, North Carolina Comprehensive Site Assessment Comments Dear Mr. Sideris: Secretary On September 8, 2015, the Division of Water Resources (Division) received the Comprehensive Site Assessment (CSA) Report for the above listed facility. A preliminary review of the CSA has identified data gaps and deficiencies that require clarification. As the Division continues its review of the assessment report, additional data gaps and deficiencies may be identified. Resolution of data gaps and deficiencies are needed in order to refine remedial design and are expected to be addressed in the CSA Supplemental Report. Please be advised the CAP Report is due on November XX, 2015 per the Coal Ash Management Act (CAMA). General Comments While the CSA Report provides data to develop a general understanding of site conditions, the report fails to fully explain groundwater exceedances of groundwater quality standards as required by CAXIA § 130A-309.211. The relationship of the distribution of constituents of interest (COIs) with respect to the coal ash basins is not well discussed or evaluated in the report. Understanding, using site data, and communicating why a contaminant is elevated in one area (or location) and low in another, is essential to assessing fate, transport, and risk to receptors prior to the development of a CAP. The CSA Report generally fails to quantitatively estimate background concentrations in soil and groundwater and use these estimates as a basis for delineating constituents. The CSA Report should provide provisional estimates of constituent concentrations using site data. The estimates are needed to compare ash leachate-impacted groundwater to background levels. 1636 Mail Service Center, Raleigh, North Carolina 27699-1636 Phone: 919-807-64641Internet: http://www.ncwater.org An Equal Opportunity 1 Affirmative Action Employer —Made in part by recycled paper NCDENRO193372 Marshall Steam Station Page 2 of 8 Specific Comments The data gaps and deficiencies noted during review of the CSA Report include the following: • The Division expects "provisional estimates for purposes of assessment (PEPA)" of background concentration levels for constituents of interest (COIs) in groundwater to be established in order to evaluate the results of the CSA Report consistent with CSA Guidelines and CAMA requirements to describe all exceedances of North Carolina Administrative Code (NCAQ Title 15A Chapter 02L.0202 groundwater quality standards (2LStandards). Although various ranges of background concentrations (regional, MW-7D/SR, and BG series wells) were provided in the report, overall site provisional estimates of background levels for the COI are needed as a basis for comparison of CSA data and determination of the nature and extent of contamination. The PEPA determinations should be made using background data from relevant compliance wells (historic and current data) as well as background data from relevant assessment wells. Data outliers influenced by turbidity should be removed prior to the PEPA. determination. The methodology used to determine the PEPA and a list of removed outliers should be included in the CSA supplement. 0 It is indicated throughout the CSA that there are constituents that are naturally - occurring. Naturally -occurring constituents are generally in equilibrium in water/rock geochernical systems. Contamination disturbs the equilibrium by increasing concentration in one of the phases and the systems reacts by transferring some of the contaminant to the other phase until equilibrium is re-established. Concentrations in solution above the 2L Standards of naturally -occurring constituents may be an indication that equilibrium between phases has been altered. This has not been noted in the CSA Report. 0 If contamination associated with the site is determined to be the ultimate cause of naturally -occurring constituents above the 2L Standards, these constituents should be considered contaminants of interest (COIs). Therefore, the Division emphasizes that background concentrations of all constituents with concentrations above the 2L standards must be determined, and that those constituents be included in the geochernical modeling for the CAP. 0 For some naturally occurring COIs, their naturally occurring levels will be the site cleanup levels. The background levels for COIs that are naturally occurring but also whose concentrations include contributions by coal ash or whose concentrations are effected by geochernical conditions related to the ash management basin should be determined in the CSA report and CAP, 0 COI determinations should consider hexavalent chromium. • Concentrations of COIs exceeding the 2L/IMAC and likely background ranges were detected in monitoring wells installed close to, or beyond, the compliance boundary on the hydraulically up -gradient (west, south, northwest) and down -gradient (east) sides of the site. Immediately downgradient of the site is a major hydraulic discharge zone. Immediately upgradient of the site is a topographic high which the report proposes restricts the flow of groundwater to the west (upgradient direction). Although groundwater flow does typically follow the Slope -Aquifer model in piedmont settings under natural flow conditions, the high concentration of residential NCDENR0193373 Marshall Steam Station Page 3 of 8 water supply wells immediately adjacent and west of the ash basins and those to the northwest, combined with the long history of hydraulic sluicing into the ash basins, may cause or have previously caused variances of groundwater flow from the traditional Slope -Aquifer model. 0 Although current groundwater flow measurements support general flow patterns to the east/southeast toward the Catawba River/Lake Norman from the topographic divides roughly Sherrills Ford Rd and Island Point Rd, the planned hydraulic modeling should include an evaluation of whether collective private water supply well pumping and previous mounding and geochemical conditions in the ash storage areas may have affected historic contaminant migration and constituent concentrations now observed in upgradient wells GWA-2S/D, MW-llS/D, MW-1.2S/D, GWA-9BR, MW-13S/D, GWA- 3S/D, GWA-4S/D, and GWA-8S/D. Additional off -site upgradient monitoring wells may be needed to fully delineate the extent of COIs that exceed the PEPA background concentrations which are proposed. 0 Some of the groundwater constituent isoconcentration maps in Figures 10-75 through 10-122 show closed boundaries delineating the extent of 2L/lMAC exceedances where no data has been collected beyond the exceedances detected in monitoring wells. While it is understood that several COI appear to have background concentrations that exceed the 2L/L'vIACs, proposed PEPA background concentrations need to be well defined to demonstrate that the COI have been fully delineated in groundwater. Areas of uncertainty with respect to isoconcentration delineation of COI on figures should be clearly designated. Any areas where delineation of exceedances has not been achieved will require additional monitoring well nest installation and assessment. If a monitor well (data point) does not exist between a known 2L (or background) exceedance and a receptor (stream or supply well, e.g.), a monitor well should be installed and sampled to fill this data gap. If there is uncertainty about local flow directions due to an insufficient number or placement of monitor wells, additional monitor wells should be installed and sampled. The horizontal extent of a contaminant plume must be based on concentration data rather than inferences about groundwater flow and contaminant concentrations downgradient; local flow directions and concentrations must be known. o Although Total Chromium concentrations at MW-1 ID are below 2L limits, compliance monitoring data show concentrations have been increasing in this location since February 2013. Vanadium was detected at 23.8 ug/L at MW-1 ID in the July, 2015 assessment sample. Hexavalent chromium was detected at 4.6 ug/L in the July, 2015 samples from MW- I I D. Additional assessment to west of MW- I I D is encouraged to develop more off -site data to characterize groundwater hydraulic and geochemical conditions. C, Exceedances of the 2L/IMAC for Cobalt, Vanadium and Iron were reported in the MW-12 well nest. Additional off -site assessment and evaluation to the west of MW-12 is encouraged. 0 Exceedances of the 2L/1MAC for Antimony, Chromium, Vanadium and Manganese were reported in the GWA-3 well nest. Additional off -site assessment and evaluation to the northwest of GWA-_3 ) is encouraged. 0 CS.A. Section 10.2 discusses the background wells BG-1 S/D, BG-2S/BR and BG-3S/D. The text indicates that there is currently insufficient data to statistically qualify the BG series wells as NCDENR0193374 Marshall Steam Station Page 4 of 8 background but that they are likely representative of upgradient conditions. MW-4 and MW-4D were historically used as upgradient monitoring well locations. 0 Previous historic data from existing background wells MW-4 and MW-4D did not include all COIs being evaluated in the CSA. (vanadium and cobalt are examples). 0 The CSA report Section 10.2 indicates an additional bedrock and deep background monitoring well will be considered in a location south of Lake Norman and east of the Marshall Station. The Department supports the installation of an additional background well nest. Samples from the background well should also include speciation. 0 CSA report Section 10.2 and Appendix G discuss prediction limits for comparing whether historical data from background well MW-4 & MW-4D are statistically valid for determining site specific background concentrations because there are currently insufficient data sets from the BG series wells to use in statistical analyses. There are insufficient data sets from MW-4S & MW-4D for vanadium, cobalt and hexavalent chromium for statistical analyses however, PEPA background concentrations will need to also be proposed for these constituents. 0 Given the number of tested private water supply wells which have been designated by NCDHHS as unsafe for potable use due to hexavalent chromium and vanadium concentrations, it is important that groundwater from the background wells also be speciated and a thorough understanding of chromium species distribution at and around the site be developed to determine if the ash basins are fully or partially influencing the concentrations of contaminants observed in the residential water supply wells. 0 While assessing the significance and extent of hexavalent chromium, please speciate groundwater samples from the additional background monitoring well nest and BG- 2S/BR, BG-3S/D, MW-13D, GWA-9BR, GWA-3S/D, GWA-8S/D, and GWA-2S/D. CSA report Section 3.0 discusses source characteristics and volumes. Of the unlined features containing coal combustion residuals (CCRs), the report does not specifically reference the volume of CCRs in the 394 acre unlined ash basin or the structural fill underneath the photovoltaic farm. The report indicates that there are 280,000 tons of CCRs in the Division of Waste Management permitted Phase I unlined landfill and 2,515,000 tons of CCRs in the unlined Phase 11 landfill. Duke Energy's website lists the volume in the ash basin as 16,020,000 tons, the volume of CCRs in "fills" as 7,360,000 tons. Based on these numbers, the combined volume of unlined CCR storage at the facility is approximately 26,175,000 tons. Please verify and list the volumes of CCRs in the unlined and lined site storage features in the supplement and CAP. • CSA Section 4.0 discusses pertinent structures located proximal to ash management features illustrated in Figure 4-3. Please show the flow directions through pipes illustrated in Figure 4-3. 1 CSA report section 4.3 discusses NCDENR's Well Water Testing Program of drinking wells adjacent to the facility. Appendix B provides tabulated sampling results provided by Duke Energy from NCDENR. The CSA does not discuss and evaluate the detections of COls found in residential water supply wells and provide explanations as to their presence and whether or not they fall within expected background ranges. Evaluation and discussion of the sample results from adjacent water supply wells relative to COls associated with the site should be NCDENR0193375 Marshall Steam Station Page 5 of 8 conducted in the CSA Supplement and CAP along with provisions for alternate water required under CAMA. • The CSA. report evaluates COIs in soil and sediment relative to the NC Preliminary Soil Remediation Goals (PSRGs) for Industrial Health and/or the Protection of Groundwater. These values come from a table maintained by the Division of Waste Management- Superfund Section - Inactive Hazardous Sites Branch (IHSB) and are used to evaluate assessment and remediation benchmarks which should be achieved in soils and sediments. The IHSB PSRGs are used by that program to regulate sites and include a component of risk based evaluation of soil contamination during assessment and remediation. The 2L rules, which apply to the permitted coal ash facilities, require assessment and remediation of soils so that they will not be a continuing source for groundwater contamination. There is no specific component related to "health based" direct contact standards in the 2L rules for evaluating soil contamination. Assessment of soil impacts should only be measured against the Protection of Groundwater goals in the referenced table and not the Industrial Health PSRGs. • It was noted that the detection limits associated with some solid phase (soil) constituents (Antimony, Arsenic, Cobalt, Selenium, Thallium) were higher than the Protection of Groundwater goals to which they were being compared in the CSA tables. It may be necessary to resample or collect additional samples. • CSA Section 7.3.1 indicates that samples collected from boring SB-9, located approximately 900 ft northwest of boring SB-8, were not analyzed due to a laboratory error. Data from this source area location should be recollected and analyzed or a justification provided as to why this data point is no longer necessary. • In various tables in the CSA which list solids analytical data (7-2 Cation/Anion Sample Results, 8-2 Total Inorganic Results -Background Soil, 8-3 Total Inorganic Results — Background and PRW, 8-4 Total Inorganic Results -Soil, 8-5 Total Inorganic Results -Rock, 9-2 Sediment Sample Results -Total) only total chromium results are available. Sections 7.1, 7.2, 7.2, 8.3, and 9.2 of the CSA also discusses comparison of the soil results to applicable levels. The PSRG Protection of Groundwater values used in the referenced IHSB table only have separate values for trivalent and hexavalent chromium due to the large differences in toxicity between these two compounds. The CSA tables use the PSRG Protection of Groundwater values for trivalent chromium as surrogate regulatory comparison for solids analytical results from the site rather than the more restrictive PSRG Protection of Groundwater table values for hexavalent chromium. For the purposes of delineating exceedances of contaminants in soils and sediments to develop a remediation plan, solids samples from the site should be speciated for chromium to evaluate whether there are exceedances of the PSRG Protection of Groundwater values for the purposes of delineation of soil and sediment contamination. The referenced IHSB PSRG table also includes a formula which can be used to calculate a value for total chromium which would be protective of the state's 2L groundwater quality standard. • Although the geochernical site conceptual model (GSCM) presented in section 13.3 provides a "bullet list" of factors that may affect leaching, sorption and desorption, and precipitation and dissolution, the list is so generic that it is of limited value and says nothing about what is being observed geochernically at the site. While the CSA includes raw contaminant concentrations, NCDENRO193376 Marshall Steam Station Page 6 of 8 selected valence state data, pH, dissolved oxygen, eH, TDS, alkalinity, and other data in various tables and figures throughout the report, these were not used or interpreted to develop a conceptual understanding of contaminant mobility and transport at CSS. The GSCM section 13.3 should provide a summary narrative of the current understanding of geochemical conditions across the site (different areas will be characterized by different conditions and should be described accordingly) and how they are expected to affect individual contaminant mobility and transport. Specifically, the GSCM narrative should describe, based on site data: a) the source of iron and manganese in groundwater (coal ash versus native soil) and how these constituents are solubilized/mobilized and (or) precipitated at the site, b) the source of each of the other COls (coal ash versus native soil), c) the mechanism(s) and conditions under which each COI would be expected to be mobilized or attenuated at the site, d) prevailing geochemical conditions (i.e. pH and eH) in COI source areas and along identified flow paths, and e) any other factors relevant to the occurrence and mobility of contaminants at the site. Information in the GSCM narrative should explain what is known about why a COI is found at elevated concentrations in one area of CSS and at very low concentrations in another. It is understood that geochemical modeling will be performed as part of the CAP and that the GSCM will undergo revisions as part of that process. • Because the mobility of iron and manganese is controlled by geochemically-mediated (pH, eH) precipitation rather than sorption onto iron hydroxides (and Kd measurements are not particularly relevant to Fe and Mn concentrations along a flow path), a separate sub -section should be provided in section 1.3 that describes how the fate and transport of these constituents will be modeled. • Although data needed to develop a hydrogeologic site conceptual model (HSCM) are provided in various tables, figures, and appendices throughout the CSA, there is not the GAP -required section that interprets this information in a summary narrative needed to evaluate CSA results of contaminant extent and movement, ensure compliance with CSA guidelines and LAMA, and assess modeling results and CAP. Section 6.2.4, Hydrogeologic Site Conceptual Model, is the section that is supposed to provide this narrative; instead it lists the types of data that comprise the HSCM and the sections in which these data may be found. Section 6.2.4 should include a stand-alone narrative for each area and its corresponding downgradient footprint for which a coal ash source has been identified. The narrative should summarize the current conceptual hydrogeologic understanding of that area of the site. Each narrative should succinctly describe: a) the character of the connected three-part groundwater system in which flow occurs and where units are particularly thick or thin, b) how, where, and how much recharge occurs in that area, c) horizontal and vertical flow directions in that area, d) the areas of discharge, e) locations in that area that do not follow the HSCM for the area and why, and f) data gaps that affect the ability to understand a) through e) above. Areas of the site that do not fit the generalized HSCM should be discussed in the narrative. • CSA report Section 2.2 refers to Figure 2-3 which depicts the site prior to construction of the ash basins on a USGS 1954 topographic map. Figure 2-3 has a note indicating that the source map is a USGS 1981 topographic map. It appears the text reference is accurate and not the note on Figure 2-3. NCDENR0193377 Marshall Steam Station Page 7 of 8 CSA report Section 2.7 discusses permitted activities and waste at the site with site features illustrated on Figure 2-4. Please illustrate the locations and flow directions of Internal Outfall 004, Yard Sump Overflow Outfalls 002A and 002B, and Non -contact cooling water Outfall 003 on Figure 2-4. Cross -Section C to C', Figures 8-5.1, 8-5.2 and 8-5.3 are incorrectly labeled as to their orientation. The transect starts with GWA-6S/D which is illustrated on Figure 8-5.1 at the west side of the transect, however, this monitoring well nest it is actually located on the northeastern side of the transect. MW-I I S/D is illustrated as the eastern end of the transect when it is actually the southwestern end. CSA report Section 9.1. discusses one surface water sample (SW-6) which was collected from an unnamed tributary to Lake Norman adjacent to the unlined dry ash landfill (Phase 1, DWM permit 1804-ES-DUS). Cobalt was detected above the 2B standards. Although SW-6 is a surface water sample, the report noted that Boron, Manganese, and Vanadium exceeded the 2L/E\4AC standards. Monitoring wells MW-145, MW-14D, AL -IS, AL -ID, located between the Phase I landfill and the unnamed tributary, also exhibited 2L/IMAC exceedances of Boron and various other metals. Additional assessment of the extent of surface water impacts to the unnamed tributary and Lake Norman should be conducted. Additional delineation of the groundwater exceedances to the north-northeast of the unnamed tributary should be conducted. An understanding of the groundwater/surface water interactions in this area should be included in the site conceptual model. CSA report Section 10.9 and Table 1.0-17 discuss radionuclide testing in limited monitoring well locations at the site. Radium 226 was detected in monitoring well MW -14D with an activity level of 96.1 J pCi/L. Please collect and evaluate additional samples for radionuclides from downgradient flow paths in this area to include monitoring wells AB -ID, AB-IBR, AL - IS, AL -ID, AB-IIS, andAB-IID. No surface water samples were collected from the Catawba River/Lake Norman as part of the CSA investigation even though exceedances of the 2B stream standards were detected in SW-6 from the unnamed tributary that flows into Lake Norman. Exceedances of the 2L standards, which also would exceed the 2B stream standards in surface water, were noted in monitoring wells located immediately adjacent to Lake Norman. Sampling and assessment of impacts to the Catawba River/Lake Norman should be included in the assessment. An understanding of the groundwater/surface water interactions (hydraulic and geochemical) should be included in the site conceptual model. The above comments were based on a preliminary review of the CSA Report. The Division is continuing to review the report and may provide additional comments in subsequent letters, if appropriate. Failure to address the deficient items stated above may result in the assessment of civil penalties and/or the use of other enforcement mechanisms available to the State. If you have any questions, please contact Bruce Parris at (704) 663-1699. Sincerely, NCDENRO193378 Marshall Steam Station Page 8 of 8 S. Jay Zimmerman, P.G., Director Division of Water Resources cc: WQROS — MRO WQROS Central Files DENR Secretary — Don van der Vaart HDR (Attn: William Miller) 440 South Church Street, Suite .1000, Charlotte, NC 28202 NCDENR0193379