Loading...
HomeMy WebLinkAboutNC0004979_Amended Expert Report_20161213RECEIVEDINCDEOUNR DEC 13 2015 t1, Perna+� I,,, weciion Amended Expert Report of Douglas J. Cosler, Ph.D., P.E. Chemical Hydrogeologist Adaptive Groundwater Solutions LLC Charlotte, North Carolina Allen Steam Station Ash Basins Belmont, North Carolina April 13, 2016 Introduction Site Background The Allen Steam Station is a five -unit, coal-fired generating station owned by Duke Energy and located on a 1,009 -acre site near Belmont in Gaston County, North Carolina, adjacent to the west bank of the Catawba River (Lake Wylie). Coal combustion residuals ("coal ash") have been disposed since 1957 in an ash basin system located south of the station, consisting of an unlined active ash basin (formed in 1973) and an inactive ash basin The ash basin system is generally located in historical depressions formed from tributaries that flowed toward the Catawba River. For example, HDR notes in CSA Section 12 2 2 15 that the southern portion of the -active ash basin was constructed over two streams The inactive ash basin consists of. two unlined dry ash storage areas (1996); two unlined structural fill units (2003-2009), and a double -lined dry ash landfill (2009), commonly referred to as the Retired Ash Basin (RAB) Ash Landfill The RAB is located on top of the inactive ash basin immediately northeast of the active ash basin The ash storage areas and structural fill units overlie the western part of the inactive ash basin Duke Energy performed biannual voluntary groundwater monitoring at the site from May 2004 to November 2010 and NPDES permit -required monitoring starting in March 2011 Groundwater sampling results at Allen 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 (LAMA) also required owners of surface impoundments containing coal combustion residuals (CCR) to conduct groundwater monitoring and assessment and prepare a CSA report The August 2015 CSA prepared by HDR Engineering, Inc. of the Carolinas (HDR) for the Allen site determined that ash handling and storage at the Allen site have impacted soil and groundwater beneath and downgradient from the ash basin system The CSA report identified Constituents of Interest (COI) considered to be associated with potential impacts to sod 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 Allen site consists of two parts CAP Part 1 (submitted to DEQ in November 2015) provides a summary of CSA findings, further evaluation and selection of CGI, a site conceptual model (SCM), the development of 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 2 Information Reviewed My opinions are based upon an analysis and technical review of (i) hydrogeologic and chemical data collected at the Allen 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 This amended report contains additional opinions based on my review of the recently -issued CAP Part 2 report 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, Allen Steam Station Ash Basin (August 23, 2015); (2) Corrective Action Plan, Part 1, Allen Steam Station Ash Basin (November 20, 2015); (3) Corrective Action Plan, Part 2, Allen Steam Station Ash Basin (February 19, 2016), (4) Miscellaneous historical groundwater and soil concentration data for the Allen 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. 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 3 facilities, nuclear waste disposal sites, hazardous waste incinerators, and various industrial facilities I have investigated the following dissolved, ronaqueous-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, radionuclid-3s, and biological constituents Summary of Opinions The following is a brief'summary of the op.mons developed in my report • A total of 44 Compliance Boundary groundwater samples exceeded North Carolina groundwater standards for these COI. boron, chromium, cobalt, iron, manganese, sulfate, total dissolved solids, and vanadium Of these 44 exceedances, 29 were greater than the proposed provisional background concentrations by HDR, which exaggerate background levels; • Most of the background wells are either likely to be downgradient from coal -ash source areas, or appear to be downgradient from coal ash, and data from these wells overestimate actual site background concentrations, • The statistical analyses of background groundwater concentrations at the Allen site (well AB -1 R) 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, • Mayor limitations of the CAP grcundwater flow and chemical transport models prevent simulation and analysis of off-site migratio-i, • 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 Conditicn or Cap -in -Place Scenario groundwater concentrations of 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 concentration reductions at the Compliance Boundary that are generally two to four times greater compared to Cap -in -Place, best reduces impacts to surface water, and reduces cleanup times by a factor of 2.5-5 Additional excavation of secondary sources would further accelerate concentration reductions; • CSA data show multiple exceedances of groundwater standards in bedrock not only at the compliance boundary but also inside the CB However, the CAP Closure Scenarios do not 4 address either concentration reduction or off-site chemical migration control in the fractured bedrock aquifer; • Due to an incorrect boundary -condition representation of the active ash basin, the CAP models underestimate by more than a factor of two both the mass loading of COI into the Catawba River and the corresponding Catawba River 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, boron, chromium, hexavalent chromium, cobalt, and vanadium), • Future Compliance Monitoring at the Allen 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 Allen Site Introduction The groundwater system at the Allen 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 ash basins overlie native soil and was constructed in historical drainage features formed from tributaries that flowed toward the Catawba River using earthen embankment dams and dikes HDR notes in CSA Section 12.2.2 15 that the southern portion of the active ash basin was constructed over two streams A large percentage 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 west to east toward the Catawba River (Lake Wylie). Vertical groundwater flow between the three layers also occurs, and surface water ponding in the active ash basin effects flow directions locally. Near the western limits of the ash basin system is a groundwater "divide" where the flow direction changes from west to east to a general westerly flow direction. The location of this divide and the nature of groundwater flow west of the ash basin system were data gaps in the CSA and CAP Part 1 investigations, however, 5 the CAP Part 2 did not further address this ssue On the eastern site boundary the CSA and CAP Parts 1 and 2 investigations assumed that all groundwater at the Allen site (overburden and bedrock aquifers) discharges into the Catawba River HoweNer, 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 tc site groundwater flow directions, or the risk of off-site migration of COI in the overburden or bedrock aquifers, caused by groundwater extraction from numerous private and public water supply wells located close to the site boundaries and near the eastern shore of the river 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 Allen 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 Throughout 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 sods 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 (sod) 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 at the Allen site) 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 Allen 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 9 overburden (soil) or fractured bedrock aquifers, groundwater flow and COI transport into or beneath the r Catawba River 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 Allen site. In this test the static water level in the open hole (boring) or well casing is 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 great as a factor of 50 smaller than the correct Kvalues I discuss the impacts of these major errors on HDR's groundwater flow and chemical transport assessments and the CAP modeling throughout the remainder of my report Bedrock Slug Tests HDR made mayor mistakes in computing the bedrock hydraulic conductivity values using the field monitoring -well slug -test results (CSA Appendix H) The Barker and Black (1983) slug test analysis method used by HDR for bedrock wells was misapplied in all cases The Barker -Black method is an analytical solution that simulates water level changes in a well that fully penetrates a fractured isotropic aquifer The Barker -Black solution does not incorporate the well screen length as a parameter, as the Bouwer-Rice (1976) method does (see following section). However, the screen length (length of the slotted, or open, portion of the monitoring well) largely controls the hydraulic head changes in a partially - penetrating (i e , monitoring) well, along with the horizontal and vertical hydraulic conductivities of the aquifer (Hantush, 1964, Bear, 1979). The ratio of the aquifer thickness to screen length is also important As an illustration, in bedrock monitoring well GWA-3BR HDR used a well screen length of 248 feet ("saturated thickness" in their Appendix H Barker -Black analysis), but the true screen length in this well is only five feet Therefore, the Barker -Black model "thinks" a 248 -foot thick aquifer is causing the water - level change during the slug test, but the actual effective saturated thickness is about a factor of 50 7 smaller (i e , on the order of the five-foot screen length) Using five feet for the saturated thickness in the GWA-3BR slug test analysis would yield a true K estimate that is about a factor of 50 higher (— 248/5) More accurate analyses of the bedrock slug tests also need to account for partial -penetration effects (i.e , the hydraulic effects of a pumping well with a "short" well screen located in a thicker aquifer). Due to these errors the bedrock slug test results presented in the CSA are on the order of a factor of 50 too low In the CAP flow model the bedrock is assumed to be factors of 40 to 4,000 lower than the overburden aquifer in different areas (CAF Appendix C, Table 2) When these errors are corrected the overall mean hydraulic conductivity for the bedrock unit is similar in magnitude to the overburden sods and transition zone (CSA Table 11-10). If the mean bedrock permeability was actually 50-100 times lower than the overburden aquifer there would be significant resistance to overlying groundwater and dissolved COI migrating into the bedrock unit and flow and transport would tend to preferentially flow in the overburden aquifer in a more predominantly horizontal direction (e g , parallel to the bedrock surface) However, that is not the case at the Allen site because the bedrock and overburden permeabdities are similar in magnitude and downward groundwater flow in the upper portion of the bedrock unit is not restricted Moreover, the Allen site conceptual hydrogeologic model, the CSA field investigation and data analyses, and the CAP flow and transport models were based on this incorrect bedrock hydraulic conductivity assessment. I discuss these issues further in the "groundwater flow" section of 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 a vertical anisotropy, A, = KnonzonratIKvertioat , that is a factor of five lower than the values presented in the CSA report (e.c , compare geometric mean values in CSA Tables 11-10 and 11- 11) and used in the CAP modeling (e.g , CAP 1 report Appendix C, Table 2) The values of A„ in the CSA and CAP modeling are generally on the order of ten for overburden soils, whereas the Bouwer-Rice slug test analyses assumed A„ = 2 (CSA Appendix H) Using the correct vertical anisotropy, A,, = 10, increases all of the measured overburden hydraulic conductivity values (CSA Table 11-4) by about 25 percent, depending on how the slug -test radius of influence was computed Since every reported overburden K value in the CSA report (at least for new shallow and deep wells) is approximately 25 percent too low, the actual average chemical transport rates in overburden soils are about 25 percent greater than reported. This site -wide data reduction error also affects the CAP 1 and 2 flow and transport model calibrations For example, HDR significantly reduced laboratory measurements of the sod -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 Kvalues had been used to 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 9 of COI with soil, the chemical migration rate is proportional to K/ Kd (except for Kd << 1) The CAP 1 transport model history matching indicated that the simulated transport rate was too low, so the model developers reduced the model Kd In other words, the reductions in calibrated Kd values would not have been as great if the correct (higher) Kvalues were used in the first place. As discussed below, 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 are still 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 COI sorption to soil is important because, as discussed below, 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 Throughout 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 Allen 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 the Catawba River and (n) groundwater flow directions and the potential for offsite migration of COI. Catawba River and the Le Grand Conceptual Model All groundwater at the Allen site was apparently assumed to discharge into the Catawba River according to a generalized conceptual model (LeGrand, 2004) before actual site-specific hydrogeologic data were analyzed. Statements to this effect 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 site groundwater to discharge into the river at the downgradient model boundary. HDR continued to state this assumption in the CAP 2 report (e.g , Section 3 3 2) even though the consistently downward groundwater flow components next to the Catawba River show that this assumption is incorrect (refer to the deep/bedrock aquifer vertical hydraulic gradients measured during the CSA and CAP Part 2 investigations) In CAP 2 Section 3.3 2, HDR also states that "The Catawba River serves as the lower hydrologic boundary for groundwater within the shallow, deep, and bedrock flow layers at the site " However, the river cannot be the "lower hydrologic boundary" for the deep and bedrock layers when the measured vertical flow direction in these layers is consistently downward, which demonstrates that HDR has not delineated this inferred "lower boundary" used in the CAP models. HDR further states in CAP 2 Section 3 3 2 that "the approximate vertical extent of the groundwater impacts is generally limited to the shallow and deep flow layers." This statement ignores that fact that groundwater flow across the site is consistently downward from the impacted deep flow layer to the highly -fractured bedrock aquifer and that, as discussed below, 12 exceedances of North Carolina 2L and/or IMAC groundwater standards (and greater than background concentrations) were measured in samples collected from bedrock wells located inside the Compliance Boundary. 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 no= 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 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 fielc 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 or stream and the depth of hydraulic influence within an underlying aquifer are highly sensitive to several factors, including. the transient river water surface elevation and slope; river bed topography, bed permeability and thickness, horizontal and vertical permeability (and thickness) of the different hydrogeologic units underlying the river, transient horizontal and vertical hydraulic head variations in groundwater beneath and near the river, 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 river bed permeability or thickness; characterize the river bathymetry; monitor transient water surface elevation variations at more than one location (one average value was used), collect river bed hydraulic gradient data, measure horizontal or vertical overburden or bedrock permeability beneath or on the east side of the river; characterize the geology beneath or east of the river, measure hydraulic heads in the overburden or bedrock beneath or east of the river, or consider the hydraulic effects of groundwater extraction from nearby water supply wells (e.g., the numerous private homes located on the east bank of the Catawba River, near the southeast corner of the site, and west of the property boundary) Much of the data that were collected in the CSA contradict the LeGrand hypothesis at the Allen site For example, strong downward flow components from deep to bedrock wells were measured at each deep/bedrock monitoring cluster (GWA-5, GWA-3, and GWA-1) next to the Catawba River (compare CSA Figures 6-6 and 6-7), and the following shallow -deep well clusters on the west river bank GWA-4, AB -9, and AB -10 (CSA Figures 6-5/6-6) Downward groundwater flow was also measured at several other locations across the site (CSA Table 11-14). Water -level measurements 10 collected in September 2015 for the CAP Part 2 investigation [CAP 2 Figures 2-2 (shallow), 2-3 (deep), and 2-4 (bedrock)] demonstrated downward groundwater flow from the deep to bedrock aquifers at all locations across the Allen site with the exception of monitoring well clusters AB-25D/BR and AB-35D/BR Maps of vertical hydraulic gradient variations (e g., contour maps) were not generated for the CSA or CAP Part 2, and HDR did not discuss the significance of downward hydraulic gradients next to the river and at most other deep/bedrock monitoring well clusters These downward groundwater flow measurements are consistent with the hydraulic conductivities of the bedrock and overburden being of similar magnitude, as discussed above The strong and consistent measured downward groundwater flow components right next to the Catawba River and at most other well clusters indicate that site groundwater is entering the deep fractured bedrock unit at the downgradient property boundary (and many other areas of the site) and that not all of the site groundwater discharges into the river as the site Conceptual Model and the CAP flow and transport models assume The downward flow into bedrock at the downgradient property boundary may also be due in part to groundwater extraction from private bedrock water supply wells located along the eastern shore of the river, but in the CSA and CAP investigations HDR assumed these factors related to the potential for off-site COI migration beneath the river were not important and did not evaluate them The assumption that all site groundwater discharges into the Catawba River may also be related to the sparse number of bedrock monitoring wells (two) that were installed along the entire downgradient property boundary. 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 As an illustration, the slug -test analysis mistakes have resulted in several inaccurate hydrogeologic interpretations and conclusions in the Allen CSA with regard to groundwater flow and chemical transport rates and directions (horizontal and vertical) Because the permeability of the weathered bedrock is similar to the overlying sods the CSA and CAP interpretations that the bedrock acts as a lower confining layer for groundwater flow and chemical transport is incorrect As further discussed below in the model review section, the much higher (after correction for errors) bedrock permeability also generally increases the potential for off-site COI migration toward private and public water supply wells Therefore, the CSA 11 and CAP conclusions that (i) all site groundwater discharges into the Catawba River and (ii) groundwater and dissolved coal -ash constituents are restricted from migrating to the west toward the residential and water supply wells are not consistent with the data Exceedances of Groundwater Standards In this section I compare measured groundwater concentrations in shallow, deep, and bedrock groundwater samples to North Carolina 2L and [MAC standards and show the following (i) 34 measured exceedances for several COI at multiple locations on the Compliance Boundary (CB); (ii) an additional 10 exceedances at CB locations beneath the Catawba River based on chemical transport modeling I performed, (iii) 29 of the 44 Compliance Boundary exceedances were greater than the proposed provisional background concentrations (PPBC) by HDR, (iv) 20 of the 44 Compliance Boundary exceedances were greater than the maximum background concentration in the same hydrogeologic unit (e.g, shallow, deep, or bedrock), (v) 13 additional exceedances were measured in wells located on the Duke Energy property boundary (PB) in areas where the CB and PB do not coincide, as drawn by HDR; (vi) 12 additional exceedances were observed in wells screened in the highly -permeable fractured bedrock unit underlying the ash basin and located inside the CB; (vii) several background wells are likely downgradient from coal ash and cannot be relied upon to provide accurate COI background levels, (viii) excluding these questionable backgrourd wells which I believe to be downgradient from coal ash, 32 of the 44 Compliance Boundary exceedances were greater than the maximum background concentration in the same hydrogeologic unit; and (ix) the statistical analyses of groundwater concentrations at well AB - 1 R for purposes of defining background levels were performed incorrectly [HDR also believes AB -1 R may be downgradient from COI discharging from the inactive ash basin (CSA Report Section 10.2); in the CAP 2 report HDR states that AB -1 R will no longer be used as a background well due to increasing COI concentration trends at this location] Throughout this report I reference the ash basin compliance boundary and the Duke Energy property boundary for the Allen site as drawn on maps developed by HDR (e g , CSA Figure 6-2) 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 the Ash Basin Compliance Boundary (CB) as drawn by HDR; (ii) on the Duke Energy property boundary (PB), where the CB and PB do not coincide; (iii) bedrock wells (BR) located inside the CB; and (iv) modeled Compliance Boundary concentrations 12 (CBM), using modeling techniques described below The proposed provisional background concentrations (PPBC) by HDR are also listed in Table 1. A total of 34 Compliance Boundary groundwater samples exceeded North Carolina groundwater standards for these COI- boron, chromium, cobalt, iron, manganese, sulfate, total dissolved solids, and vanadium I estimated an additional 10 exceedances at CB locations beneath the Catawba River based on chemical transport modeling and measured upgradient concentrations (CBM) In addition, 12 exceedances were observed in wells screened in the highly fractured bedrock unit located inside the CB Thirteen more exceedances were measured in wells located on the Property Boundary in areas where the CB and PB do not coincide, according to maps provided with the CSA and CAP. Note that the CAP 2 transport model also predicts year 2015 Compliance Boundary exceedances for these COI* antimony [shallow(S), deep(D), bedrock(BR)]; boron (S,D), chromium (S,D,BR), hexavalent chromium (S,D,BR), cobalt (S,D,BR); sulfate (S,D,BR); and vanadium (S,D,BR) (CAP 2 report, Table 4-1) Excluding questionable background wells which may be downgradient from coal ash, (footnote "g" in Table 1; refer to discussion below), 32 of the 44 (measured plus modeled) Compliance Boundary exceedances were greater than background levels (from the same hydrogeologic unit) for a particular constituent. Nine of the 13 PB exceedances were greater than background levels Because the bedrock background well BG-2BR may be downgradient from coal ash, it appears that all of the 12 bedrock exceedances were greater than BR background levels Ignoring the potential problems with background well (BG) locations (i.e , using all background wells), 20 of the 44 Compliance Boundary exceedances were greater than any BG concentration (from the same hydrogeologic unit) for a particular constituent. Five of the 13 PB exceedances were greater than maximum background levels The concentrations for 10 of the 12 bedrock exceedances were greater than any BR background level. A total of 29 of the 44 Compliance Boundary exceedances were greater than the proposed provisional background concentrations (PPBC) by HDR, which exaggerate background levels due to their reliance on data from monitoring wells that I believe are downgradient from coal ash Seven of the 13 property boundary exceedances were greater than the PPBC. Note that the iso -concentration contours in all of the CSA Section 10 figures and Figure ES -3 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 the downgradient Duke Energy property boundary which coincides 13 with the western shore of the Catawba River Figure ES -3 is a good example of this practice. These closed contours at the downgradient property boundary suggest that COI transport beyond the farthest downgradient line of monitoring wells does not occur and that no COI reach the Compliance Boundary located beneath the river (where there are no monitoring wells) This is not the case, of course, for several COI as demonstrated in the following section, where COI exceedances at the Catawba River section of the Compliance Boundary are demonstrated by modeling, and illustrated by most of the CAP model simulated "existing conditions" plume maps in CSA Appendix C which contain 'open contours" at the river shoreline. If the CAP model grid did not stop short of the Compliance Boundary (CB) the Appendix C existing condition COI plumes would reach the CB in most cases Modeled Compliance Boundary Exceedances I computed Compliance Boundary (CB) concentrations labeled "CBM" with footnote "e" in Table 1 using a calibrated one-dimensional, analytical chemical transport model (van Genuchten and Alves, 1982, Equation C5) because the CB at these locations was up to 500 feet offshore beneath the Catawba River. The CAP transport model grid did not extend to this downgradient portion of the CB, and HDR did not evaluate COI concentrations at this locat,on I calibrated the analytical model to chemical -specific site conditions (i e., determined model input parameter values) using CAP transport model simulated concentration versus time curves for "Existing Conditions" (CAP report Appendix C) The analytical model input parameters in my model were: groundwater pore velocity, chemical retardation factor, and longitudinal dispersiwty Figures 1 a and 1 b show the close agreement between the analytical model simulation results (solid lines) for Boron (Well GWA-4S) and Cobalt (Well GWA-5S), respectively, and the CAP model predictions from Appendix C (solid circles) For each constituent, I used the calibrated analytical model to compute the concentration versus time curve immediately downgradient at the Compliance Boundary (dashed lines) The modeled CBM concentrations in Table 1 are equal to the product of the measured monitoring well concentration and the ratio of the simulated CB and simulated monitoring well concentrations for 2015 (Figure 1). Exceedances of Groundwater and Surface Water Standards in Seep Samples Concentrations in several seep water samples exceeded relevant NCAC 213, 2L and/or IMAC standards for various COI (e g , CSA Table 7-8, CSA Appendix F, CSA Figure 2-2, CAP Table 2-5.1 and 2-5 2) As discussed in the introduction, the active ash basin at the Allen site is constructed above two historical streams that are tributaries of the Catawba River (e.g., CSA Figure 2-3). Seeps S-3 and S-4, located downgradient of the ash basin between the toe of dams and the Catawba River, coincide with the downgradient portions of the historical streams visible on topographic maps (CAP Figure 2-2). Recent surveys identified streams at these locations and at locations coinciding with seep locations S-1 and S-2, 14 as well as a wetland complex (CAP Figure 1-5). HDR describes seep locations S-1 through S-4 as tributaries flowing toward the river (Lake Wylie) with well-defined stream channels of varying widths (2014 Seep Water and Monitoring Plan, Table 1). Referring to my Table 1, 14 of the seep samples exceeded North Carolina groundwater standards (CSA Table 7-8) for these COI boron, iron, manganese, and vanadium. These samples are from seep locations S-3, S-4, ANWW-002, ANWW-004, ANSW-001, and ANSW-015 The CAP report describes seep locations S-1 and S-2 as "dry" and does not report sampling results, although other reports describe streams with continuous flow at these locations. North Carolina surface water (26) standards have been exceeded in seeps samples for these constituents (locations not specified): aluminum, copper, lead, manganese, mercury, and total dissolved solids (CAP Table 2-5 2). CSA Appendix F notes that previous seep sampling at locations S-1 through S-9 detected 213 exceedances for iron, manganese, zinc, and thallium Determination of Background Concentrations Background Well Locations Most of the background wells at the Allen site have not been installed in locations that are clearly upgradient from site -related coal -ash materials For example, designated background well ABA R is likely downgradient from the Inactive Ash Basin and Ash Storage Areas based on the hydraulic head measurements at wells AB -38S and ABA R (e g , the 622-, 623-, and 624 -foot contours) In this regard, Section 10 2 of the CSA report discusses the fact that a recent trend of increasing coal -ash constituent concentrations has been observed at this location and acknowledges that AB -1 R may be downgradient from the Inactive Ash Basin The CSA report further recommends that data from this well needs to be carefully evaluated to determine whether it can continue to be used for background monitoring In the CAP 2 report HDR states that ABA R will no longer be used as a background well due to these problems In addition, well 13G-1 S is directly downgradient from the western half of the Active Ash Basin based on the 636- and 637 -foot hydraulic head contours in CSA Figure 6-5 (HDR did not draw these specific contours) The light -blue "approximate groundwater flow direction" arrows in this area are incorrectly drawn Well BG -2S may be downgradient from the Inactive Ash Basin and the Ash Storage Areas according to hydraulic head measurements in these areas Deep background well BG -1 D is likely downgradient from the western half of the Active Ash Basin based on hydraulic head measurements in this area [e g , the 636 -foot hydraulic head contour in CSA Figure 6-6 (not drawn by HDR)] The light -blue "approximate groundwater flow direction" arrows in this area are 15 incorrectly drawn. Well BG -2D may be downgradient from the Inactive Ash Basin and Ash Storage Areas. Bedrock background well BG-2BR may be downgradient from site -related coal -ash materials based on the bedrock hydraulic head map (CSA Figure 6-7) and due to groundwater extraction from private and public water supply wells located immediately west of the Allen site As discussed throughout my report, the CSA and CAP modeling failed to analyze the effects of off-site groundwater extraction. Statistical Analyses of Background Concentrations Appendix G of the CSA report presents statistical analyses of historical concentrations from shallow Monitoring Well AB -1 R, 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 Allen 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 statisticady 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 AB -1 R background data 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 at the Allen site, can lead to the same general volume of the chemical plume being repeatedly sampled when the monitoring events are closely spaced. For example, the groundwater pore velocity (Vp) near well AB -1 R is on the order of 8 feet per year based on the measured hydraulic conductivity (slug test results in CSA Table 11-5) and the horizontal hydraulic gradients in this area (CSA Figure 6-5) Note that shallow pore velocities are generally a factor of 200 greater in many areas downgradient of the ash basin system due to much greater hydraulic gradients (- 20x larger) and larger hydraulic conductivity(— 10x greater) in this area In addition, groundwater pore velocities in deep overburden and in fractured bedrock are generally more than a factor of 1,000 greater than velocities in the shallow overburden (CSA Table 11-13) 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) Accordingly, based on measured sorption values and average permeability values for the aquifer the mean chemical migration rate (Vp /Rd) near background well ABA R is on the order of 0.1 ft/yr for several of the non -conservative COI, assuming linear equilibrium sorption (refer to discussion below) For the ABA R quarterly sampling, the ncn-conservative chemical migration distance between ABA R 16 sampling rounds (3 months) is less than 0.1 feet for some COI, which is approximately equal to the sandpack diameter for the monitoring wells. Therefore, the ABA R background samples are basically representative of the same volume of the plume (i e , the vicinity of the sandpack, depending on the well purge volume) for many COI, and the measured sample concentration changes analyzed in the CSA are not due to actual chemical transport effects in the overburden aquifer. This means that the groundwater samples are non -independent and that the statistical analyses of background concentrations at ABA R 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 Parts 1 and 2 groundwater flow models I focus specifically on model boundary conditions representing the Catawba River; 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. Catawba River Boundary Condition The CAP Parts 1 and 2 groundwater flow models force all groundwater at the downgradient property boundary to discharge directly into the Catawba River and significantly underestimate the potential for off- site flow and chemical migration in fractured bedrock [The Electric Power Research Institute (EPRI) model review letter (October 20, 2015) refers to the use of drain (i e , 'leaky type") boundary conditions in the flow model, through which 17 percent of the site groundwater discharges from the model, but the CAP reports do not show their location or clearly discuss their hydraulic effect on the flow regime.] No -flow boundary conditions defined along the entire western, northern, and southern model boundaries and the bottom two layers (i e , bedrock) of the eastern boundary prevent any off-site flow and chemical transport in these areas (refer to Figures 1 and 4 in Appendix C of the CAP 1 Report) The bottom surface (bedrock) of the flow model is also assumed to be a no -flow boundary even though the hydraulic conductivity data and measured downward hydraulic gradients at most deep/bedrock monitoring well clusters contradict this assumption The only locations where groundwater and dissolved constituents are allowed to leave the CAP models (other than the unexplained drain leakage at the top of the model) are the vertical array of cells located above the bedrock unit along the eastern grid boundary, which are assumed to represent the Catawba River, these cells are specified as constant -head boundary conditions in which the head is uniform with depth. 17 This hydraulic representation of the Catawba River in the flow models is inaccurate for many reasons First, the river 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 river 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 river, the CAP models do not represent actual site hydrologic conditions Groundwater flow at the Allen site is not strictly horizontal and, as discussed above, most of the deep/bedrock vertical hydraulic gradient measurements (including next to the river) are downward Third, as represented in the CAP models, neither the lower -permeability river bed sediments nor the smaller vertically hydraulic conductivity of underlying sods restricts the potential flow rate into or out of the river (i.e , in the CAP models a perfect hydraulic connection exists between the aquifer and the Catawba River) The actual degree of aquifer -river hydraulic connection was not evaluated in the CSA or CAP Parts 1/2 In summary, due to all of these factors the potential for site groundwater and dissolved constituents to migrate off-site eastward beyond the Catawba River or southward as underflow beneath the river cannot be evaluated with the CAP models. As I discuss in this report, the consistent measured downward groundwater flow components from deep overburden into the highly -permeable bedrock aquifer next to the Catawba River and at most other locations across the site indicate that COI may be migrating off-site beneath the river, which is a mayor contradiction of the Allen site conceptual model The CAP models should have represented the Catawba River using a 'leaky -type" (i.e., river) boundary condition in the top model layer (McDonald and Harbaugh, 1988), and the model grids should have extended farther east so that HDR coulc have evaluated the above factors during model calibration and sensitivity analyses In their reviews of both the CAP 1 and 2 models (submitted with the CAP modeling appendices), the Electric Power Research Institute third -party peer review team also concluded that the Catawba River should be modeled as a leaky boundary condition instead of using constant heads The models also should have included groundwater extraction from the private water supply wells installed at many points close to the eastern river bank. A river boundary condition incorporates the bed permeability and thickness, the river water surface elevation, and the simulated hydraulic head in the aquifer (at the base of the 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 river and aquifer. Typically, permeability and vertical hydraulic gradient measurements for the river bed (not collected in the CSA) and flow model calibration (three-dimensional matching of simulated and measured hydraulic head measurements in the aquifer) are used to determine a best -fit estimate of river bed conductance (permeability divided by thickness) in the model HDR did not perform this routine analysis 18 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 to include the South Fork Catawba River and also incorporated groundwater extraction from off-site private -home and public water -supply wells The western no -flow boundary in the current CAP 1 and 2 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 despite the fact that several private homes are also located south of the Active Ash Basin and shallow, deep, and bedrock hydraulic head maps exhibit southeasterly flow components in this area Artificial limitations created by the eastern Catawba River boundary condition are outlined above In addition, the downgradient boundaries of the CAP 1 and 2 flow and transport models do not extend to the Compliance Boundary (except for a small interval near Well GWA-3), 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 of the same order of magnitude as the overburden soils based on my corrections for slug test analysis errors. In the present configuration the lower boundaries of the CAP Parts 1 and 2 model grids are only about 50 feet below the bedrock surface (Figure 3 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 As discussed above, the strong measured downward hydraulic gradients between deep and bedrock wells in both the June and September 2015 sampling rounds (e g , Wells GWA-5, GWA-3, GWA-1 next to the Catawba River, and Well BG -2) demonstrate that vertical and horizontal groundwater flow in bedrock is important, and these transport mechanisms need to be accurately simulated in the CAP models in order to accurately assess the potential for off-site chemical migration. Model Significantly Underestimates Leakage Rate from Active Ash Basin The CAP Part 2 flow model underestimates leachate discharge from the active ash basin by as much as a factor of 100 in areas of ponded surface water (e g., refer to CSA Figure 4-3) The CAP 1 model underestimates active basin leakage by as much as a factor of 200 As shown in Figure 5 of CAP 1, Appendix C, the CAP 1 flow models assumes a constant groundwater recharge rate (i.e., basin leakage rate) equal to 8.5 inches/year in the active ash basin In the CAP 2 flow model the active basin leakage rate is assumed to be 14 5 inches/year (Figure 5 of CAP 2, Appendix B) However, CSA Figure 8-8.2 (cross-section F -F) shows that the vertical hydraulic gradient through the coal ash in the downgradient portion of the active ash basin is on the order of unity Using Darcy's law and the mean vertical coal -ash 19 permeability of 1 6E-4 cm/sec in CSA Table 11-11, the approximate vertical leakage rate out of the active basin is about 2,000 inches/year near the Catawba River (i e., — 140 times greater than the specified recharge rate of 14 5 inches/year) The CAP flow models should have represented ponded areas of the active ash basin as either constant - head or leaky -type boundary conditions, which would have allowed the model to simulate a realistic leakage rate for the active ash basin. The mayor discrepancies between the measured shallow hydraulic head maps (CSA Figure 6-5 and CAP 2 Figure 2-2) and the CAP 1 and 2 simulated shallow head maps (Figure 14 in CAP 1, Appendix C, Figure 14 in CAP 2, Appendix B) clearly show that the CAP flow models significantly underestimate the hyd-aulic head beneath the active ash basin due to the fact that the modeled leakage rate from the active Lasin is much too low. Three related impacts of this incorrect active basin boundary condition are that the CAP models significantly underestimate (i) vertical groundwater flow rates (by on the order of a factor of 100) through coal -ash source material in the vicinity of the downgradient portion of the active ash basin, (u) horizontal groundwater flow and chemical transport rates (— factor of two) downgradient from the active ash basin (compare measured and simulated horizontal hydraulic gradients between the ash basin and the Catawba River); 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 u-iderestimate (by more than a factor of two) both the mass loading of COI into the Catawba River and the corresponding Catawba River 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) 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 Allen site to migrate with groundwater to private and public bedrock water supply wells located immediately west, south, and east (near the eastern bank of the Catawba River) of the Duke Energy property boundary CSA Figure 4-2 shows the locations of private and public water supply wells near the site The basis of my opinion includes the following hydraulic conductivity measurements for the overburden and bedrock formations, including my corrections to the bedrock slug test analyses, three- dimensional variations in measured hydraulic head in the bedrock and overburden units, groundwater concentration data, and calculations of potential hydraulic head reductions (i.e , drawdown) that could be caused by off-site groundwater extraction As discussed throughout my report, neither the CSA nor CAP Parts 1 or 2 investigations addressed the potential for off-site migration. 20 COI 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 Allen site 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 Part 1 modeling did not include these areas west of the Allen site. The CAP Part 2 flow model did include a small number (44 of the 219 neighboring private wells and none of the four public water supply wells) of residential wells located inside the undersized model domain (between Southpoint Road and the active ash basin), 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 and water supply wells located immediately west, south, and east of the site (eastern shoreline of the Catawba River), fix the boundary condition problems, or correct the model input data errors so that the flow and transport models could be used to more accurately analyze the potential for off-site chemical transport. One of the mayor input data errors is the bedrock hydraulic conductivity, which is assumed in the CAP Parts 1 and 2 flow models to be factors of 100 to 10,000 lower than the overburden aquifer in most areas (CAP 1 Appendix C, Table 2, CAP 2 Appendix B, Table 2) A corrected analysis of the bedrock slug test data shows that the mean bedrock permeability is approximately the same as the overburden permeability. Therefore, downward groundwater flow from the overburden aquifer into the upper portion of the bedrock unit is not restricted, as the CAP flow models represent, and the potential for off-site chemical migration is underestimated by the models 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 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 Allen site because for most of the COI the bulk of the chemical mass in the sod is associated with the solid phase (i.e., sorbed to soil grains 21 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 + pb Kd In, 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 accirate Kd values in the CAP Closure Scenario modeling Specifically, the CAP Part 1 transport modeling used Kd values that are typically factors of 10 - 100 (i e , one to two orders of magnitude) smaller than the measured site-specific Kd's reported in CAP 1 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 are still 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 Note that the Kd values in CAP 1 Appendix D are also simaar in magnitude to EPA measurements at the Allen Site (EPA, 1985, Table 5 4) Further, sod -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 Allen site, the times required to reach North Carolina water quality standards at the Compliance Boundary are at least a factor of 10 longer (see additional discussion below) than cleanup times predicted by the CAP Parts 1 and 2 transport models for many COI. The CAP 1 modeling report (CAP 1 Appendix C; Section 4 8) argues that the major Kd reductions were needed due to the following (a similar statement is made in Section 4 8 of CAP 2 Appendix B). "The conceptual transport model specifies that COls enter the model from the shallow saturated source zones in the ash basins. When the measured Kd values are applied in the numerical model to COls migrating from the source zones, some COls do not reach the downgradient observation wells where they were observed in JunelJuly 2015 at the and of the 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. These may include pond dredging, dewatering for dike construction, and ash grading and placement. This 22 approach is expected to produce conservative results, as sorbed constituent mass is released and transported downgradient." Considering the approach that was 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 Parts 1 and 2 transport models used an incorrect value (2.65 g/cm3) for the bulk density of overburden materials; this value is the density of a solid mass of mineral (e.g , quartz) with zero porosity. The bulk density should have been computed using the total porosity (n) values in CSA Table 11-1 using the following formula (e.g., Baes and Sharp, 1983). pb = 2.65 (1— n) Based on the Table 11-1 values pb — 1 1 - 1 8 g/cm3, which means that the Rd values for the CAP 1 and 2 models were as much as a factor of 2 4 (2 65/1 1) too high before HDR adjusted the Kd values during calibration Also, as discussed earlier, the overburden slug test values were about 25 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 up to three times 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 (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 Allen 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 (To-a„ei ) for a constituent to reach a downgradient observation point, and HDR used differences in observed versus simulated Ttrave/ (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 the source was activated 58 years ago and that: "a source term matching the pore water concentrations for each COI was applied within the inactive ash basin, active ash basin and the ash storage areas at the start of the calibration period. The source 23 concentrations were adjusted to match measured values in the downgradient monitoring wells that had exceedances of the 2L Standard for each COI in June 2015." For several reasons it is a major simplification (and generally inaccurate) to use 2015 ash pore water concentrations to define year -1957 source zone (fixed concentration) boundary conditions. These reasons include coal ash was gradually and nonuniformly distributed (spatially and temporally) in ash basins throughout the 58 -year simulation period (not instantaneously in 1957), 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 1957 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 Allen site and are only qualita=ive 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 condiFons 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 (u) "Quantitative predictions of chemical attenuation rates based upon mineralogy and groundwater composition cannot be made because only descriptive and qualitative information are available for adsorption/desorption mechanisms " Nonetheless, HDR performed the geochemical modeling to evaluate the technical basis for its MNA analysis, however, any quantitative MNA analysis must compare mass transport rates and changes (e g., 24 grams/year per unit area normal to a groundwater pathlme) 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 Allen 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 "Data collected to date show that COls will attenuate over time to restore groundwater quality at the Allen site." HDR also claims in CAP 2, Section 6 3 3 that "the groundwater model did not allow for removal of COI via co -precipitation with iron oxides, which likely resulted in an over -prediction of COI transport. Completion of the Tier II assessment described in Appendix H has addressed this issue." 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 fads 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 sod 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, boron, chromium, hexavalent chromium, cobalt, 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 16) were that antimony, boron, cadmium, chromium, cobalt, manganese, and nickel show limited evidence of attenuation and should not be evaluated further for MNA These Appendix H conclusions are consistent with the Table 4-1 results for these COI In addition, my Table 1 results support the Appendix H conclusions that boron, chromium, cobalt, and manganese are not candidates for MNA because these contaminant plumes have also already reached 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 Allen site " 25 Simulation of Closure Scenarios As discussed below, CAP 1 Closure Scenaro 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 (CIP) remedial alternative evaluated in the CAP Part 1, the Excavation Scenario results in COI concentration reductions at the Compliance Boundary that are generally two to four times (2 - 4x) greater compared to Cap -in -Place and best reduces impacts to surface water In addition, the time frames to achieve equivalent concentration reductions are factors of 2 5 to 5 shorter for excavation compared to cap -in-place. 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) in the CAP 2 model due to the much larger Kd (and Rd) values 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 Scenano In this scenario the CAP 1 flow model predicts a cap -induced water -table decline equal to approximately 26 feet (relative to the Existing Conditions simulation) in the center of the inactive ash basin (24 feet in CAP 2 flow modeling), the predicted water -table decline in the active ash basin is about 36 feet (32 feet in CAP 2) The geologic cross-sections presented in the CSA show that the saturated coal ash thickness at several locations is up to twice the value assumed in the Cap -in -Place simulation For example, the CSA geologic cross-sections show saturated coal ash thicknesses in the active ash basin equal to about 54 feet at boring AB -25 and in cross-section E -E , 95 feet across cross-section B -B ; 47 feet at boring AB -21, and 70 feet across section F -F Saturated coal ash thicknesses in the inactive ash basin are equal to approximately 43 feet at boring AB -29, 29 feet at boring AB -29, and 29 feet in cross-section D- D. 3E This means that under the simulated Cap -In -Place Scenario up to one-half 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 It is important to note, however, that the CAP groundwater flow model simulations exaggerate the hydraulic effects of the cap (i e , overstates water table lowering) due to the no -flow boundaries and the incorrect (i.e., significantly too low) bedrock permeability 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 flow model uses a hydraulic conductivity value that is on the order of 50-100 (or greater) times too small, thus restricting upward flow from bedrock into the capped area and exaggerating predicted water table lowering 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 CAP 1 flow model uses an assumed uniform value of 5 inches year outside of the Ash Basin System and 8.5 inches/year inside ash basins (except zero recharge for the Retired Ash Basin) even though the actual values are highly variable across the Ash Basin System and site land surface Further, as discussed above, HDR should have used a leaky -type boundary condition to model ponded areas of the active ash basin. The predicted water table lowering due to capping is very sensitive to the model recharge value, so more effort should have been made to develop a site-specific recharge -rate distribution Slow and Multirate Nonequdibnum 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 remediation system is turned off (Sudicky and Illman, 2011; Hadley and Newell, 2014; Culver et al., 1997). 27 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 Scenano simulations did not utilize this modeling feature Slow desorption of COI can also be expected at the Allen 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 (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 model solves 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 sensitivity analyses that were performed only varied Kd by +/- 20 percent (Table 7 in CAP 1 Appendix C), which is much too small to address the actual uncertainty 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 MT3D 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 Allen 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 As outlined in my report, the CAP Part 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 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 mayor errors in model input data, model calibration mistakes, field 28 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 factors of 10 or more errors in cleanup time predictions with the CAP 1 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 CAP 1 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 is insufficient to dewater all of the source -area coal ash, as discussed above, and the CAP 1 and 2 flow models overestimate cap -induced water -table lowering due to boundary condition errors. Furthermore, these simulation times are well beyond the prediction capabilities of any chemical transport model for a complex field site (especially one that is as geochemically complex as the Allen site). The historical model -calibration dataset (1957-2015) 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 Although the CAP 1 model underestimates remediation time frames, the CAP 1 Closure Scenario simulations demonstrate several significant advantages of excavation for restoring site groundwater quality versus cap -in-place First, predicted COI concentration reductions in groundwater downgradient from the ash basin system are generally factors of 2-4 greater with excavation compared to CIP (e g., refer to most of the simulated concentration versus time curves in CAP 1 Appendix C). Further, if HDR had correctly performed the cap -in-place simulations the predicted CIP concentrations would be much higher because predicted water -table lowering due to the cap would be insufficient to dewater all of the coal ash Second, North Carolina 2L or IMAC standards for many COI are not achieved by CIP but are achieved by excavation (e.g., CAP 1 Appendix C, Figures 16, 17, 31, 74, and 104-107). Third, the time frames to achieve equivalent concentration reductions are factors of 2 5 to 5 (2 5 - 5x) shorter for excavation compared to cap -in-place [e g , Antimony, 2 5x (Fig 17), 4x (Fig 16); Arsenic, 5x (Fig 30), Barium, 5x (Fig 44), Chromium, 4x (Fig 74); Cobalt, 3x (Fig. 105); Selenium, 2 5 (Fig 121) ] Even though 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-half of the saturated coal -ash 29 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 geochem►cal modeling indicates that attenuation by a combination of sorption, chemical precipitation, and dilution effectively dissipates COls in groundwater in the vicinity of the ash basins" and that "... it is reasonable to assume that CO/s remaining in groundwater will decrease in concentration over time as upgradlent 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 one-half 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 HDR also incorrectly concluded in the recommendations section of CAP 2 Appendix G that "... there are only two locations where groundwater that has been reported to exceed 2L standards fingers out to the compliance boundary" I showed above (and in Table 1) that groundwater standards have been exceeded at multiple locations around the Compliance Boundary As discussed above, the CAP Part 2 flow model did include a small number of residential wells (44 of the 219 neighboring private wells and none of the four public water supply wells) located between Southpoint Road and the active ash basin, 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 capture COI dissolved in groundwater 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 16 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 1957 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 west, south, north, and east the capture zones for the remaining 175 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 not only at the compliance boundary but also inside the CB In addition, strong downward groundwater flow components from the deep overburden to bedrock aquifers were measured during the CSA at multiple locations 30 across the site (and at all but two deep/bedrock well clusters in CAP 2), including the western shoreline of the Catawba River. The cap -in-place alternative does not address either concentration reduction or off- site chemical migration control in the fractured bedrock aquifer. The CAP Parts 1 and 2 do not assess whether water quality standards will be achieved in the tributaries and wetlands between the ash basins and the Catawba River under any closure scenario Tributaries downgradient of the active ash basin currently receive contaminated groundwater discharge from the active ash basin at seep locations S-3 and S-4 As discussed above, for the cap -in-place scenario a significant fraction of the source material will remain saturated and dissolved COI will continue to migrate with groundwater toward these seep locations. Although unaddressed by the model, COI concentration decreases in groundwater and unsaturated zone pore water due to source removal would also reduce impacts to tributaries and wetlands that are influenced by the ash basins Conclusions Based on my technical review and analyses of the referenced information for the Allen site I have reached the following conclusions. • A total of 44 Compliance Boundary groundwater samples exceeded North Carolina groundwater standards for these COI. boron, chromium, cobalt, iron, manganese, sulfate, total dissolved solids, and vanadium Of these 44 exceedances, 29 were greater than the proposed provisional background concentrations by HDR, which exaggerate background levels; • Most of the background wells are either likely to be downgradient from coal -ash source areas, or appear to be downgradient from coal ash. As a result, groundwater concentration data from these wells overestimate actual site background levels, • The statistical analyses of background groundwater concentrations at the Allen site (well AB -1 R) are invalid The time penods between groundwater sample collection from this well are too small and the concentration data are not independent, • 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 (Parts 1 and 2) domain size, the no -flow boundary conditions along the western, southern, and northern boundaries; the incorrect hydraulic conductivity value for the bedrock aquifer, and incorrect hydraulic boundary condition representations of the Catawba River and the active ash basin prevent simulation and analysis of off-site COI migration, 31 • The CAP Closure Scenario simulations greatly underestimate (by factors of 10 or more) the time frames required to achieve meaning;ul groundwater concentration reductions in response to remedial actions. This is due to oversimplification of field fate and transport mechanisms in the CAP model and several model input errors; • The simulated water table lowering for the Cap -in -Place Scenario is almost a factor of two too low at several locations in the ash basin system in order to dewater all source material; and the actual cap -induced water table elevation reduction would be much less than predicted due to the incorrect no -flow boundary conditions and bedrock permeability in the flow model Therefore, the remediation time frames for this scenario would be much greater because part 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 coal -ash constituents much higher t-ian 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 concentration reductions at the Compliance Boundary that are generally two to four times greater compared to Cap -in -Place and best reduces impacts to surface water In addition, the time frames to achieve equivalent concentration reductions are factors of 2 5 to 5 shorter for excavation compared to cap -in-place, and source removal reduces the number of COI that will exceed North Carolina groundwater standards in the future 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 more than a factor of two both the mass loading of COI into the Catawba River and the corresponding Catawba Rider 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, demonstrated that MNA is not an effective remedial option for several COI (e g , antimony, boron, chromium, hexavalent chromium, cobalt, 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 multiple exceedances of groundwater standards in bedrock not only at the compliance boundary but also inside the CB. The cap -in-place alternative does not address either concentration reduction or off-site chemical migration control in the fractured bedrock aquifer, and 32 • 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 33 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 Sod Using Distributed Mass Transfer Rates Environmental Science and Technology, 31(6), 1581-1588. Culver, T.B , R A Brown, and J.A Smith 200 Rate -Limited Sorption and Desorption of 1,2 - Dichlorobenzene to a Natural Sand Soil Column. Environmental Science and Technology, 34(12), 2446-2452. 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 Nqhoff, Dordrecht, Netherlands Gelhar, L W 1986 Stochastic subsurface hydrology from theory to applications. Water Resources Research 22 135S -145S Gelhar, L W. 1987 Stochastic analysis cf 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 N'hoff, 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 CJ. Newell 2012 Groundwater Remediation The Next 30 Years Groundwater, Vol 50, No 5 669-678 34 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 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. Molz, F 2015 Advection, Dispersion, and Confusion. Groundwater, Vol 53, No 3 348-353. 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 Siostrom, 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 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.0 , 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. 35 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. 36 1400 1200 J C 1000 C O w 800 c v C O V 600 c 8 O m 400 200 0 �p r r rr as a a a a 1 CAN M.W a GWA4S a - e Comphancs Dwndary a a a a a a A a 1 l a d P a J a tOD O00 O N V (00 NO O N � N N N N N N N N Time (years) 35 30 ,j 25 5 a} s r r r 04LLf I I 0 0 0 0 0p 0 0 0 0 0 .0 O O O f0 M O N Q .— N N N N N N N N Time (years) Figure 1 Modeled Compliance Boundary Concentrations 37 Case: 15-17447, 05/31/2016, ID: 9997388, DktEntry: 40, Page 1 of 45 No. 15-17447 IN THE UNITED STATES COURT OF APPEALS FOR THE NINTH CIRCUIT HAWAII WILDLIFE FUND; SIERRA CLUB -MAUI GROUP; SURFRIDER FOUNDATION; WEST MAUI PRESERVATION ASSOCIATION, Plaintiffs -Appellees, V. COUNTY OF MAUI, Defendant -Appellant. RECEWEDIN DeojDvVp DEC 3 X0'6 On Appeal from the U.S. District Court, Dist. of Hawaii No. 12-cv-198, Hon. Susan Oki Mollway, District Judge BRIEF FOR THE UNITED STATES AS AMICUS CURIAE IN SUPPORT OF PLAINTIFFS APPELLEES OF COUNSEL: KARYN WENDELOWSKI U.S. Environmental Protection Agency Office of General Counsel Washington, D.C. JOHN C. CRUDEN Assistant Attorney General AARON P. AVILA R. JUSTIN SMITH FREDERICK H. TURNER Attorneys, U.S. Dep't of Justice Env't & Natural Resources Div. P.O. Box 7415 Washington, DC 20044 (202) 305-0641 frederick.turner@usdoj.gov Case: 15-17447, 05/31/2016, ID: 9997388, DktEntry: 40, Page 2 of 45 TABLE OF CONTENTS TABLE OF AUTHORITIES.................................................................... iii INTEREST OF THE UNITED STATES..................................................1 ISSUESPRESENTED..............................................................................2 STATEMENT OF THE CASE.................................................................. 3 I. STATUTORY BACKGROUND................................................................ 3 II. FACTUAL BACKGROUND................................................................... 6 III. PROCEDURAL BACKGROUND............................................................. 7 SUMMARY OF ARGUMENT.................................................................10 ARGUMENT...........................................................................................13 I. THE DISTRICT COURT'S DECISIONS ARE CONSISTENT WITH THE LANGUAGE AND PURPOSE OF THE CWA...................................13 A. Discharges of Pollutants to Jurisdictional Surface Waters Through Groundwater with a Direct Hydrological Connection Properly Require CWA Permits ...........................14 B. The District Court's Decisions Give Full Effect to Congress's Intent to Restore and Maintain the Nation's Waters.......................................................................20 C. The District Court's Finding of Liability is Consistent with EPA's Longstanding Position..........................................22 II. THE COUNTY IS LIABLE FOR UNPERMITTED DISCHARGES DUE TO THE "DIRECT HYDROLOGICAL CONNECTION" BETWEEN THE GROUNDWATER AND THE OCEAN ............................................. 26 III. THE DISTRICT COURT CORRECTLY HELD THAT THE COUNTY HAD FAIR NOTICE FOR PURPOSES OF CIVIL PENALTIES .................. 32 CONCLUSION.......................................................................................36 i Case: 15-17447, 05/31/2016, ID. 9997388, DktEntry, 40, Page 3 of 45 CERTIFICATE OF COMPLIANCE........................................................37 CERTIFICATE OF SERVICE.................................................................38 ii Case, 15-17447, 05/31/2015, ID, 9997388, DktEntry: 40, Page 4 of 45 TABLE OF AUTHORITIES Cases Bath Petrol. Storage, Inc. v. Sovas, 309 F. Supp. 2d 357 (N.D.N.Y. 2004) ............................................... 22 Chevron, U.S.A., Inc. v. NRDC, Inc., 467 U.S. 837 (1984).................................................................... 12,24 Friends of Sakonnet v. Dutra, 738 F. Supp. 623 (D.R.I. 1990)......................................................... 15 Greater Yellowstone Coal. v. Larson, 641 F. Supp. 2d 1120 (D. Idaho 2009) ........................................ 31,32 Haw. Wildlife Fund v. Cty. of Maui, No. 12-198, 2015 WL 328227 (D. Haw. Jan. 23, 2015) .... 6, 7, 8, 9, 28 Haw. Wildlife Fund v. Cty. of Maui, No. 12-198, 2015 WL 3903918 (D. Haw. June 25, 2015) ................... 9 Hawaii Wildlife Fund v. County of Maui, 24 F. Supp. 3d 980 (D. Haw. 2014) .......................................... passim Headwaters, Inc. v. Talent Irrigation Dist., 243 F.3d 526 (9th Cir. 2001).............................................................. 5 Hernandez v. Esso Std. Oil Co., 599 F. Supp. 2d 175 (D.P.R. 2009) ................................................... 19 Hudson R. Fishermen's Ass'n v. City of New York, 751 F. Supp. 1088 (S.D.N.Y. 1990) .................................................. 22 Idaho Rural Council v. Bosma. 143 F. Supp. 2d 1169 (D. Idaho 2001) ............................ 11, 18, 19, 21 Inland Steel v. EPA, 901 F.2d 1419 (7th Cir. 1990).......................................................... 22 iii Case: 15-17447, 05/31/2016, ID: 9997388, DktEntry: 40, Page 5 of 45 In re EPA & Dept of Def. Final Rule, 803 F.3d 804 (6th Cir. 2015)............................................................. 24 McClellan Ecological Seepage Situation v. Cheney, No. 86-475, 20 Envtl. L. Rep. 20,877 (E.D. Cal. Apr. 30, 1990) ....... 31 McClellan Ecological Seepage Situation v. Cheney, 763 F. Supp. 431 (E.D. Cal. 1989) .................................................... 31 McClellan Ecological Seepage Situation v. Weinberger, 707 F. Supp. 1182 (E.D. Cal. 1988) .................................................. 30 N. Cal. River Watch v. City of Healdsburg, 496 F.3d 993 (9th Cir. 2007).............................................................. 8 N. Cal. River Watch v. Mercer Fraser Co., No. 04-4620, 2005 WL 2122052 (N.D. Cal. Sept. 1, 2005) ... 16, 17, 19 Nw. Envtl. Def. Ctr. v. Grabhorn, No. 08-548, 2009 WL 3672895 (D. Or. Oct. 30, 2009) ...................... 19 O'Leary v. Moyer's Landfill, Inc., 523 F. Supp. 642 (E.D. Pa. 1981) ..................................................... 15 Rapanos v. United States, 547 U.S. 715 (2006) ...................................................... 2, 8, 10, 15, 16 Rice v. Harken Expl. Co., 250 F.3d 264 (5th Cir. 2001) ...................................................... 19,20 S.F. Herring Assn v. Pac. Gas & Elec. Co., 81 F. Supp. 3d 847 (N.D. Cal. 2015) ................................................ 18 Sierra Club v. Abston Constr. Co., 620 F.2d 41 (5th Cir. 1980) .................................................. 10, 14, 15 Sierra Club v. El Paso Gold Mines, Inc., 421 F.3d 1133 (10th Cir. 2005) ........................................................ 16 iv Case: 15-17447, 05/31/2016, ID: 9997388, DktEntry: 40, Page 6 of 45 Sierra Club v. Va. Elec. & Power Co., No. 15-112, 2015 WL 6830301 (E.D. Va. Nov. 6, 2015) ................... 18 United States v. Approximately 64,695 Pounds of Shark Fins, 520 F.3d 976 (9th Cir. 2008)............................................................ 33 United States v. Riverside Bayview Homes, Inc., 474 U.S. 121 (1985).......................................................................... 20 United States v. Velsicol Chem. Corp., 438 F. Supp. 945 (W.D. Tenn. 1976) ................................................ 16 Vill. of Oconomowoc Lake v. Dayton Hudson Corp., 24 F.3d 962 (7th Cir. 1994).............................................................. 19 Wash. Wilderness Coal. v. Hecla Mining Co., 870 F. Supp. 983 (E.D. Wash. 1994) ................................................ 21 Yadkin Riverkeeper v. Duke Energy Carolinas, LLC, No. 14-753, 2015 WL 6157706 (M.D.N.C. Oct. 20, 2015) ................ 18 Statutes 33 U.S.C. § 1251(a).................................................................................. 3 33 U.S.C. § 1311............................................................................. 3, 4, 14 33 U.S.C. § 1318(a)(A)........................................................................... 34 33 U.S.C. § 1319....................................................................................... 4 33 U.S.C. § 1319(d)............................................................................ 5,35 33 U.S.C. § 1341(a)................................................................................ 35 33 U.S.C. § 1341(a)(1)............................................................................ 31 33 U.S.C. § 1342.............................................................................. 1, 3, 4 33 U.S.C. § 1342(a)................................................................................... 4 v Case, 15-17447, 05/31/2016, ID: 9997388, DktEntry: 40, Page 7 of 45 33 U.S.C. § 1342(b).................................................................................. 4 33 U.S.C. § 1342(d).................................................................................. 4 33 U.S.C. § 1344.................................................................................. 3,4 33 U.S.C. § 1362...................................................................................... 3 33 U.S.C. § 1362(6).................................................................................. 3 33 U.S.C. § 1362(7).............................................................................. 2,4 33 U.S.C. § 1362(8).................................................................................. 2 33 U.S.C. § 1362(12)(A) ..................................................................... 3, 14 33 U.S.C. § 1362(14)................................................................................ 4 33 U.S.C. § 1365...................................................................................... 4 Federal Register 39 Fed. Reg. 43,759 (Dec. 18, 1974) ........................................................ 4 55 Fed. Reg. 47,990 (Dec. 2, 1990) ........................................................ 23 56 Fed. Reg. 64,876 (Dec. 12, 1991) .................................................. 5,23 66 Fed. Reg. 2960 (Jan. 12, 2001) ....................................... 12, 23, 24, 26 80 Fed. Reg. 37,054 (June 29, 2015) ............................................... 17,25 vi Case: 15-17447, 05/31/2015, ID: 9997388, DktEntry: 40, Page 8 of 45 The United States respectfully submits this brief as amicus curiae pursuant to Federal Rule of Appellate Procedure 29(a). INTEREST OF THE UNITED STATES The United States Environmental Protection Agency (EPA) implements the Clean Water Act (CWA), 33 U.S.C. §§ 1251-1387, together with the states. That includes promulgating regulations regarding the CWA's National Pollutant Discharge Elimination System (NPDES). Id. § 1342. The United States participates as amicus curiae because it has an interest in the proper interpretation of the NPDES- permit provisions and the framework for analyzing whether discharges of pollutants to jurisdictional surface waters through groundwater are subject to those provisions.' The United States also has an interest because it enforces the CWA and because it is a potential defendant in actions alleging the discharge of pollutants from federal facilities through groundwater. The United States agrees with the result the district court reached in this case and urges affirmance. In the United States' view, a NPDES 1 We use the term "jurisdictional surface waters" throughout this brief to mean "waters of the United States." 1 Case: 15-17447, 05/31/2016, ID: 9997388, DktEntry: 40, Page 9 of 45 permit is required here because the discharges from the Defendant - Appellant County of Maui's wastewater treatment facility are from a point source (i.e., the injection wells) to waters of the United States (i.e., the Pacific Ocean2). To be clear, the United States does not contend that groundwater is a point source, nor does the United States contend that groundwater is a water of the United States regulated by the Clean Water Act. Moreover, the United States does not agree with the district court's application of the "significant nexus" standard from Rapanos v. United States, 547 U.S. 715 (2006). ISSUES PRESENTED This amicus brief addresses the following issues: 1. Whether a discharge of pollutants from a point source to jurisdictional surface waters through groundwater with a direct hydrological connection to jurisdictional surface waters is regulated under the CWA. 2. Whether the site-specific facts here give rise to a "discharge of a pollutant" under the CWA. 2 More specifically, into the Pacific Ocean that is part of the United States' territorial seas under the CWA. 33 U.S.C. § 1362(7), (8). 2 Case- 15-17447, 05/31/2016, ID: 9997388, DktEntry: 40, Page 10 of 45 3. Whether the County had fair notice that it was subject to civil penalties for its discharges to jurisdictional surface waters without a NPDES permit. STATEMENT OF THE CASE I. STATUTORY BACKGROUND Congress enacted the Clean Water Act to "restore and maintain the chemical, physical, and biological integrity of the Nation's waters." 33 U.S.C. § 1251(a). Congress therefore prohibited any non -excepted "discharge of any pollutant" to "navigable waters" unless it is authorized by a permit. Id. §§ 1311, 1342, 1344, 1362. The CWA defines "discharge of a pollutant" as "any addition of any pollutant to navigable waters from any point source." Id. § 1362(12)(A) (emphasis added). Pollutant means "dredged spoil, solid waste, incinerator, sewage, garbage, sewage sludge, munitions, chemical wastes, biological materials, radioactive materials, heat, wrecked or discarded equipment, rock, sand, cellar dirt and industrial, municipal, and agricultural waste discharged into water." Id. § 1362(6). The CWA defines "navigable waters" as "the waters of the United States, including the territorial seas"; and a point source is "any discernible, confined and discrete M Case: 15-17447, 05/31/2016, ID: 9997388, DktEntry: 40, Page 11 of 45 conveyance, including but not limited to any pipe, ditch, channel, tunnel, conduit, well, discrete fissure, container, rolling stock, concentrated animal feeding operation, or vessel or other floating craft, from which pollutants are or may be discharged." Id. § 1362(7), (14). The CWA authorizes EPA to issue NPDES permits under Section 402(a), but EPA may authorize a state to administer its own NPDES program if EPA determines that it meets the statutory criteria. Id. § 1342(a), (b). When a state receives such authorization, EPA retains oversight and enforcement authorities. Id. §§ 1319, 1342(d). Hawaii obtained such permitting authority in 1974. See 39 Fed. Reg. 43,759 (Dec. 18, 1974). The CWA is a strict -liability regime that prohibits non -excepted discharges unless they are authorized by a CWA permit. Id. §§ 1311, 1342, 1344. An unpermitted discharge constitutes a violation of the CWA regardless of fault and is subject to enforcement by the state or federal government or a private citizen. Id. §§ 1319, 1365. To establish liability for a violation of the permit requirement, a plaintiff must show there was (1) a discharge (2) of a pollutant (3) to navigable waters (4) n Case: 15-17447, 05/31/2016, ID: 9997388, DktEntry: 40, Page 12 of 45 from a point source. Headwaters, Inc. u. Talent Irrigation Dist., 243 F.3d 526, 532 (9th Cir. 2001). The CWA includes a civil -penalty provision for those who violate the Act. 33 U.S.C. § 1319(d). When determining a civil -penalty amount, courts must consider "the seriousness of the violation or violations, the economic benefit (if any) resulting from the violation, any history of such violations, any good -faith efforts to comply with the applicable requirements, the economic impact of the penalty on the violator, and such other matters as justice may require." Id. EPA's longstanding position is that a discharge from a point source to jurisdictional surface waters that moves through groundwater with a direct hydrological connection comes under the purview of the CWA's permitting requirements. E.g., Amendments to the Water Quality Standards Regulations that Pertain to Standards on Indian Reservations, 56 Fed. Reg. 64,876, 64,982 (Dec. 12, 1991) ("[T]he affected ground waters are not considered `waters of the United States' but discharges to them are regulated because such discharges are effectively discharges to the directly connected surface waters."). 5 Case: 15-17447, 05/31/2016, ID: 9997388, DktEntry: 40, Page 13 of 45 II. FACTUAL BACKGROUND The County operates the Lahaina Wastewater Reclamation Facility. Haw. Wildlife Fund v. Cty. of Maui, 24 F. Supp. 3d 980, 983 (D. Haw. 2014) [Hawaii 1]. The facility receives approximately four million gallons of sewage each day. Id. After treating the sewage, the facility releases three to five million gallons of effluent into four on-site injection wells. Id. at 983-84. The effluent travels into a shallow groundwater aquifer and then flows into the Pacific Ocean through the seafloor at points known as "submarine springs." Id. at 984; see also Haw. Wildlife Fund v. Cty. of Maui, No. 12-198, 2015 WL 328227, at *1 (D. Haw. Jan. 23, 2015) [Hawaii H]. EPA, the Hawaii Department of Health (DOH), and others conducted a tracer -dye study that confirmed this conclusion for injection wells 3 and 4. Hawaii I, 24 F. Supp. 3d at 984. According to the study, it took the leading edge of the dye 84 days to go from wells 3 and 4 to the ocean and about 64% of the dye injected into these wells was discharged from the submarine springs to the Pacific Ocean. Id. The dye's appearance in the ocean "conclusively demonstrated that a hydrogeologic connection exists." Id. at 985-86. 0 Case, 15-17447, 05/31/2016, ID: 9997388, DktEntry. 40, Page 14 of 45 Although tracer dye was not placed into well 1 and dye from well 2 was not detected in the study, the County "acknowledge [d] that there is a hydrogeologic connection between wells 1 and 2 and the ocean." Hawaii 11, 2015 WL 328227, at *1-2. The tracer -dye study models indicated that, in some circumstances, treated effluent from well 2 would move along flowpaths similar to those traveled by the dye injected into wells 3 and 4 and emerge at the same springs. Supplemental Excerpts of Record (SER) 237, 240, 243. There is no dispute that given the proximity of wells 1 and 2, the modeling for well 2 predicts the flowpaths for discharges from well 1. Excerpts of Record (ER) 443; SER 189. III. PROCEDURAL BACKGROUND In April 2012, Plaintiffs Appellees Hawaii Wildlife Fund, Sierra Club -Maui Group, Surfrider Foundation, and West Maui Preservation Association filed suit seeking to require the County to obtain and comply with a NPDES permit and to pay civil penalties. Hawaii 1, 24 F. Supp. 3d at 986. The district court issued three partial summary - judgment opinions in favor of Plaintiffs. The parties then entered into a settlement agreement, in which the County stipulated to terms 7 Case, 15-17447, 05/31/2016, ID: 9997388, DktEntry: 40, Page 15 of 45 contingent on a final judgment that the County violated the CWA and that the County was "not immune from" civil penalties. Haw. Wildlife Fund v. Cty. of Maui, No. 12-198, ECF No. 259. The court entered final judgment in accordance with its opinions and the settlement agreement. The district court's first opinion held the County liable under the CWA for unpermitted discharges from wells 3 and 4. Hawaii I, 24 F. Supp. 3d at 1000. The court started its analysis with the language and purpose of the CWA, and also relied on EPA's interpretation and case law. Id. at 995-96. The court explained that Plaintiffs "must show that pollutants can be directly traced from the injection wells to the ocean such that the discharge at the LWRF is a de facto discharge into the ocean." Id. at 998 (emphasis in original). The court found that Plaintiffs had met this burden. Id. at 998-1000. The district court also found CWA liability under the "significant nexus" standard from Justice Kennedy's concurring opinion in Rapanos, 547 U.S. at 755-56, and the Ninth Circuit's application of that standard in Northern California River Watch v. City of Healdsburg, 496 F.3d 993, 999-1000 (9th Cir. 2007). E:3 Case: 15-17447, 05/31/2016, ID: 9997388, DktEntry: 40, Page 16 of 45 The district court's second opinion held the County liable for unpermitted discharges from wells 1 and 2. Hawaii II, 2015 WL 328227, at *6. The County "expressly conced[ed] that pollutants introduced by the County into wells 1 and 2 were making their way to the ocean," and the court rejected the County's argument that liability does not arise unless a pollutant passes through "a series of sequential point sources." Id. at *2-4. The district court's third opinion rejected the County's argument that it was not subject to civil penalties for its unpermitted discharges because it lacked fair notice. Haw. Wildlife Fund v. C,ty. of Maui, No. 12-198, 2015 WL 3903918, at *6 (D. Haw. June 25, 2015) [Hawaii III]. The court determined that the County had notice because the discharges "clearly implicate[d] each statutory element." Id. at *4. The court further held that its adjudication of the first motion for partial summary -judgment provided notice to the County. Id. at *6. The parties then entered into a settlement agreement, in which the County stipulated that it would make good faith efforts to obtain and comply with a NPDES permit and that it would pay $100,000 in civil penalties and $2.5 million for a supplemental environmental E Case, 15-17447, 05/31/2016, ID 9997388, DktEntry, 40, Page 17 of 45 project, all contingent on a final judgment and ruling that the County violated the CWA and that the County was "not immune from" civil penalties. Haw. Wildlife Fund v. Cty. of Maui, No. 12-198, ECF No. 259. The district court then entered a final judgment. SUMMARY OF ARGUMENT The judgment should be affirmed because it is consistent with the language and purpose of the Clean Water Act and EPA's longstanding interpretation and practice of issuing NPDES permits for discharges of pollutants similar to the ones here. As Justice Scalia said in Rapanos, the statute's language prohibiting "any addition of any pollutant to navigable waters from any point source" does not limit liability only to discharges of pollutants directly to navigable waters. See Rapanos, 547 U.S. at 743 (plurality op.) (emphasis in original). Courts have interpreted the CWA as covering not only discharges of pollutants directly to navigable waters, but also discharges of pollutants that travel from a point source to navigable waters over the surface of the ground or through underground means. E.g., Sierra Club v. Abston Constr. Co., 620 F.2d 41, 44-45 (5th Cir. 1980). The discharges in this case fall squarely within the statutory language. 10 Case: 15-17447, 05/31/2016, ID: 9997388, DktEntry: 40, Page 18 of 45 In the United States' view, a NPDES permit is required here because the discharges at issue are from a point source (i.e., the injection wells) to waters of the United States (i.e., the Pacific Ocean's coastal waters). To be clear, the United States views groundwater as neither a point source nor a water of the United States regulated by the CWA. The United States therefore agrees with the district court's conclusion that a NPDES permit was required here, but only to the extent that the court's analysis is consistent with the above -stated principles regarding groundwater. The district court's conclusions accord with the CWA's purpose. Congress enacted the CWA "to restore and maintain ... the country's waters"; and to achieve this goal, Congress created a strict -liability regime prohibiting discharges unless they are authorized under the CWA. Recognizing Congress's goals in the CWA, courts have concluded that in certain circumstances discharges of pollutants that reach navigable waters through groundwater fall squarely within the statute's terms. E.g., Idaho Rural Council v. Bosma, 143 F. Supp. 2d 1169, 1179-80 (D. Idaho 2001). 11 Case: 15-17447, 05/31/2016, ID: 9997388, DktEntry: 40, Page 19 of 45 Even if Congress's intent on this issue had been ambiguous, EPA has clearly stated for decades that pollutants that move through groundwater can constitute discharges subject to the CWA, and that interpretation is entitled to Chevron deference. Chevron, U.S.A., Inc. v. Nat. Res. Def. Council, Inc., 467 U.S. 837, 842-43 (1984). It has been EPA's longstanding position that discharges moving through groundwater to a jurisdictional surface water are subject to CWA permitting requirements if there is a "direct hydrological connection" between the groundwater and the surface water. See NPDES Permit Regulation and Effluent Limitations Guidelines and Standards for Concentrated Animal Feeding Operations, 66 Fed. Reg. 2960, 3017 (Jan. 12, 2001). This formulation recognizes that some hydrological connections are too circuitous and attenuated to come under the CWA. Id. The County argues that the district court dispensed with the requirements that a discharge be "from a point source" and "to navigable water" because the effluent was discharged from a nonpoint source and because the effluent was discharged into groundwater, which is not covered by the CWA. Opening Brief (Op. Br.) at 21, 27, 30. 12 Case. 15-17447, 05/31/2016, ID. 9997388, DktEntry: 40, Page 20 of 45 This attempt to bifurcate the movement of the pollutants into two separate events is inconsistent with the statute's language and purpose. It also ignores the undisputed fact that the pollutants moved through that groundwater to the ocean. The County's argument that no civil penalty should have been imposed because the County lacked fair notice lacks merit. The County was on notice both as a general matter—through the CWA's language and EPA's statements in rulemakings—and specifically—through communications from EPA to the County. In any event, the question of fair notice goes to the amount of the civil penalty, an amount the County stipulated to, and is only one of many factors informing a civil - penalty amount. ARGUMENT I. THE DISTRICT COURT'S DECISIONS ARE CONSISTENT WITH THE LANGUAGE AND PURPOSE OF THE CWA. The district court's judgment holding the County liable under the CWA is consistent with the text and purpose of the statute. It is also consistent with EPA's long -held position governing when the CWA requires permits for discharges of pollutants that move to jurisdictional surface waters through groundwater with a direct hydrological 13 Case: 15-17447, 05/31/2016, ID: 9997388, DktEntry: 40, Page 21 of 45 connection. The County cannot recast the nature of the discharges to avoid that result. A. Discharges of Pollutants to Jurisdictional Surface Waters Through Groundwater with a Direct Hydrological Connection Properly Require CWA Permits. When Congress prohibited the unpermitted "discharge of any pollutant," it defined this term broadly as "any addition of any pollutant to navigable waters from any point source." 33 U.S.C. §§ 1311, i 1362(12)(A). As the County concedes, "a point source does not need to discharge directly into navigable waters to trigger NPDES permitting." Op. Br. at 27. Because Congress did not limit the term "discharges of pollutants" to only direct discharges to navigable waters, discharges through groundwater may fall within the purview of the CWA. This reading of "discharge of a pollutant" has been applied in other similar contexts where discharges of pollutants have moved from a point source to navigable waters over the surface of the ground or by some other means. In Sierra Club u. Abston Construction, which addressed discharges from mining operations that traveled to navigable waters in part through surface runoff, the Fifth Circuit stated that "[g]ravity flow, resulting in a discharge into navigable body of water, Case, 15-17447, 05/31/2016, ID, 9997388, DktEntry: 40, Page 22 of 45 may be part of a point source discharge if the [discharger] at least initially collected and channeled the water and other materials."3 620 F.2d at 44-45; see also Friends of Sakonnet v. Dutra, 738 F. Supp. 623, 628, 630 (D.R.I. 1990) (defendant liable for discharge of "raw sewage [that] was running directly from the leaching field, on the surface of the ground for approximately 250 feet, into the [surface water]"); O'Leary v. Moyer's Landfill, Inc., 523 F. Supp. 642, 647 (E.D. Pa. 1981) ("[T]here is no requirement that the point source need be directly adjacent to the waters it pollutes."). That Congress gave the term "discharge of a pollutant" a broad meaning finds support in cases where CWA liability attached for discharges from point sources that traveled through other means before reaching surface waters. See Rapanos, 547 U.S. at 743 (noting that courts have found violations of Section 301 "even if the pollutants discharged from a point source do not emit `directly into' covered 3 The County misconstrues the United States' position as amicus curiae in Abston Construction. See Op. Br. at 30-31. The United States took the position that discharges of pollutants that traveled indirectly from a point source to jurisdictional surface waters through surface runoff or the gravity flow of rainwater come within the scope of the CWA. Brief for the United States as Amicus Curiae, at 35-36, Sierra Club v. Abston Constr. Co., No. 77-2530 (5th Cir. 1980). 15 Case: 15-17447, 05/31/2016, ID: 9997388, DktEntry: 40, Page 23 of 45 waters, but pass `through conveyances' in between") (citing Sierra Club v. El Paso Gold Mines, Inc., 421 F.3d 1133, 1137 (10th Cir. 2005) (defendant could be liable for discharges conveyed from its point -source mine shaft to jurisdictional surface water through a tunnel that defendant did not own); United States v. Velsicol Chem. Corp., 438 F. - Supp. 945, 946-47 (W.D. Tenn. 1976) (holding that CWA covered pollutants discharged from defendant's point source to jurisdictional surface waters conveyed through a sewer system that the defendant did not own)). Because courts have interpreted the term "discharge of a pollutant" to cover discharges over the ground and through other means, exempting discharges through groundwater could lead to absurd results. As one court noted, "it would hardly make sense for the CWA to encompass a polluter who discharges pollutants via a pipe running from the factory directly to the riverbank, but not a polluter who dumps the same pollutants into a man-made settling basin some distance short of the river and then allows the pollutants to seep into the river via the groundwater." N. Cal. River Watch v. Mercer Fraser Co., No. 04-4620, 2005 WL 2122052, at *2 (N.D. Cal. Sept. 1, 2005). 16 Case: 15-17447, 05/31/2016, ID: 9997388, DktEntry: 40, Page 24 of 45 The County concedes that discharges need not be direct and that a discharge through a conveyance requires a permit. Op. Br. at 27. The County argues, however, that the conveyance itself must be a point source and that because groundwater is not a point source, the district court "impermissibly 'transform [s] a nonpoint source into a point source."' Id. at 27-28, 33. The County's interpretation is flawed. Contrary to the County's argument, the district court did not eliminate the requirement that a discharge be "from a point source." All it said was that pollutants from a point source need not be emitted directly into covered waters. The case law does not require the means by which the pollutant discharged from a point source reaches a water of the United States to be a point source. While the County's statement that the statutory definition of "navigable waters" does not include groundwater is accurate, Op. Br. at 21, it is beside the point. There is no dispute that groundwater itself is not a "navigable water," 80 Fed. Reg. 37,054, 37,055 (June 29, 2015), but the district court's decisions hinge on the movement of pollutants to jurisdictional surface waters through groundwater with a direct 17 Case, 15-17447, 05/31/2016, ID: 9997388, DktEntry: 40, Page 25 of 45 hydrological connection. Such an addition of pollutants to navigable waters falls squarely within the CWA's scope. The County relies on the treatment of groundwater in legislative history, Op. Br. at 21-23, but this "only supports the unremarkable proposition with which all courts agree—that the CWA does not regulate `isolated/nontributary groundwater' which has no [effect] on surface water." Bosma, 143 F. Supp. 2d at 1180. It does not undermine the conclusion that discharges of pollutants through groundwater to jurisdictional surface waters are subject to the NPDES program. The County contends that case law does not support the district court's interpretation, Op. Br. at 35-37, but this argument largely ignores the majority of the courts that have addressed this,issue and concluded that discharges that move from a point source to jurisdictional surface waters via groundwater with a hydrological connection are subject to regulation under the CWA. See, e.g., Sierra Club v. Va. Elec. & Power Co., No. 15-112, 2015 WL 6830301 (E.D. Va. Nov. 6, 2015); Yadkin Riverkeeper v. Duke Energy Carolinas, LLC, No. 14-753, 2015 WL 6157706 (M.D.N.C. Oct. 20, 2015); S.F. Herring Assn v. Pac. Gas & Elec. Co., 81 F. Supp. 3d 847 (N.D. Cal. 2015); Hernandez Case: 15-17447, 05/31/2016, ID: 9997388, DktEntry: 40, Page 26 of 45 v. Esso Std. Oil Co., 599 F. Supp. 2d 175 (D.P.R. 2009); Nw. Envtl. Def. Ctr. v. Grabhorn, No. 08-548, 2009 WL 3672895 (D. Or. Oct. 30, 2009); Mercer Fraser, 2005 WL 2122052; Bosma, 143 F. Supp. 2d 1169. The County's reliance on other case law (Op. Br. at 35-36) is unavailing for three reasons. First, none of the cases are controlling precedent. Second, most of these decisions are inapposite because they do not address the issue of discharges of pollutants that move through groundwater to jurisdictional surface waters. In Village of Oconomowoc Lake v. Dayton Hudson, Corp., the court examined whether groundwater itself was a navigable water, i.e., a water within the meaning of the CWA. 24 F.3d 962, 965 (7th Cir. 1994). That is distinct from whether a CWA permit is required when pollutants travel to jurisdictional surface waters through groundwater with a direct hydrological connection. Third, these cases do not foreclose application of the CWA where a direct hydrological connection Bio jurisdictional surface waters can be found. In Rice v. Harken Exploration Co., the court concluded that a discharge of oil that might reach navigable waters by gradual, natural seepage was not the equivalent of a discharge to navigable waters. 250 19 Case, 15-17447, 05/31/2016, ID: 9997388, DktEntry: 40, Page 27 of 45 F.3d 264, 271 (5th Cir. 2001). The court suggested, however, that it would be open to finding a discharge had occurred through groundwater when it underscored the plaintiffs' failure to provide any "evidence of a close, direct and proximate link between [the defendant's] discharges of oil and any resulting actual, identifiable oil contamination of a particular body of natural surface water." Id. at 272. B. The District Court's Decisions Give Full Effect to Congress's Intent to Restore and Maintain the Nation's Waters. Congress's purpose in enacting the CWA—to "restore and maintain the chemical, physical, and biological integrity of the Nation's waters"—embraced a "broad, systemic view ... of water quality." United States v. Riverside Bayview Homes, Inc., 474 U.S. 121, 132 (1985). The County attempts to minimalize that goal. Adopting the County's theory would allow dischargers to avoid responsibility simply by discharging pollutants from a point source into jurisdictional surface waters through any means that was not direct. Courts have viewed the CWA's broad purpose of protecting the quality of navigable waters as a clear congressional signal that "any pollutant which enters such waters, whether directly or through 20 Case: 15-17447, 05/31/2016, ID: 9997388, DktEntry: 40, Page 28 of 45 groundwater, is subject to regulation by NPDES permit." Wash. Wilderness Coal. v. Hecla Mining Co., 870 F. Supp. 983, 990 (E.D. Wash. 1994). "Stated even more simply, whether pollution is introduced by a visible, above -ground conduit or enters the surface water through the aquifer matters little to the fish, waterfowl, and recreational users which are affected by the degradation to our nation's rivers and streams." Bosma, 143 F. Supp. 2d at 1179-80. The state's authority to protect groundwater is in no way impaired by subjecting point sources to NPDES-permit requirements to protect surface waters. Thus, the County's argument that it should not be liable here because "preservation of states' authority over the regulation of groundwater" is a "co -equal" goal of the CWA misses the mark. Op. Br. at 34-35. This emphatically is not a case about the regulation of groundwater. Instead it is about the regulation of discharges of pollutants to waters of the United States. To the extent the County's argument relies on the regulatory scheme governing disposal into wells, Op. Br. at 24-27, that is flawed because the regulation of wells under the Safe Drinking Water Act's (SDWA) Underground Injection Control (UIC) program does not preclude or displace regulation under the 21 Case: 15-17447, 05/31/2016, ID: 9997388, DktEntry: 40, Page 29 of 45 CWXs NPDES program.4 See Hudson R. Fishermen's Assn v. City of New York, 751 F. Supp. 1088, 1100 (S.D.N.Y. 1990), aff'd, 940 F.2d 649 (2d Cir. 1991) (objectives of the CWA and the SDWA are not "mutually exclusive"); see also Bath Petrol. Storage, Inc. v. Sovas, 309 F. Supp. 2d 357, 369 (N.D.N.Y. 2004). C. The District Court's Finding of Liability Is Consistent with EPA's Longstanding Position. EPA's longstanding position has been that point -source discharges of pollutants moving through groundwater to a jurisdictional surface water are subject to CWA permitting requirements if there is a "direct hydrological connection" between the groundwater and the surface water. EPA has repeatedly articulated this view in multiple rulemaking preambles. In 1990, EPA stated that "this rulemaking only addresses discharges to water of the United States, consequently discharges to ground waters are not covered by this rulemaking (unless there is a 4 The County misconstrues EPA's position in Inland Steel v. EPA, 901 F.2d 1419 (7th Cir. 1990). EPA argued that not all disposals into injection wells are discharges of pollutants under the CWA, and that the connection between the wells and navigable waters in that case was too attenuated to bring the discharges under the purview of the CWA. Id. at 1422-23. That position (embraced by the Seventh Circuit) does not mean that "injection into wells is not a discharge of pollutants requiring a NPDES permit." Op. Br. at 27. M Case: 15-17447, 05/31/2016, ID: 9997388, DktEntry: 40, Page 30 of 45 hydrological connection between the ground water and a nearby surface water body)." NPDES Permit Application Regulations for Storm Water Discharges, 55 Fed. Reg. 47,990, 47,997 (Dec. 2, 1990). And in the preamble to its final rule addressing water quality standards on Indian lands, EPA stated: [T]he Act requires NPDES permits for discharges to groundwater where there is a direct hydrological connection between groundwaters and surface waters. In these situations, the affected groundwaters are not considered "waters of the United States" but discharges to them are regulated because such discharges are effectively discharges to the directly connected surface waters. 56 Fed. Reg. at 64,982. In 2001, EPA reiterated its position: "As a legal and factual matter, EPA has made a determination that, in general, collected or channeled pollutants conveyed to surface waters via ground water can constitute a discharge subject to the Clean Water Act." 66 Fed. Reg. at 3017. EPA recognized that the determination was "a factual inquiry, like all point source determinations," adding: The time and distance by which a point source discharge is connected to surface waters via hydrologically connected surface waters will be affected by many site specific factors, such as geology, flow, and slope. Therefore, EPA is not proposing to establish any specific criteria beyond confining 23 Case: 15-17447, 05/31/2016, ID: 9997388, DktEntry: 40, Page 31 of 45 the scope of the regulation to discharges to surface water via a "direct" hydrological connection. Id. A general hydrological connection between all groundwater and surface waters is insufficient; there must be evidence showing a direct hydrological connection between specific groundwater and specific surface waters. Id. To the extent there is statutory ambiguity about whether the CWA applies to discharges to jurisdictional surface waters through groundwater, EPA's interpretation is entitled to Chevron deference. Chevron, 467 U.S. at 842-43. The County's contention that the direct -hydrological -connection standard is at odds with EPA's recently -stated position on whether groundwater is a jurisdictional water misinterprets EPA's statements. Op. Br. at 38-39. The Clean Water Rule, which was promulgated in June 2015 (and stayed by the Sixth Circuit pending further order of the court, see In re EPA & Dept of Def. Final Rule, 803 F.3d 804, 809 (6th Cir. 2015)), expressly excludes groundwater from the definition of "waters of the United States." 80 Fed. Reg. 37,054. But, as EPA clarified, the fact that groundwater itself is not jurisdictional under the CWA does not mean that pollutants that reach waters of the United 24 Case. 15-17447, 05/31/2016, ID: 9997388, DktEntry. 40, Page 32 of 45 States through groundwater do not require CWA permits. "EPA agrees that the agency has a longstanding and consistent interpretation that the Clean Water Act may cover discharges of pollutants from point sources to surface water that occur via ground water that has a direct hydrologic connection to the surface water. Nothing in this rule changes or affects that longstanding interpretation, including the exclusion of groundwater from the definition of `waters of the United States."' See EPA, Response to Comments — Topic 10 Legal Analysis (June 30, 2015); available at http://www.epa.gov/cleanwaterrule/response-comments- clean-water-rule - definition -waters -unite d- states. The County erroneously attempts to conflate the jurisdictional exclusion of groundwater with the role that groundwater can play as the pathway through which pollutants from a point source reach jurisdictional surface waters.5 5 The district court stated that if the proposed Clean Water Rule was finalized, it "would likely mean that the groundwater under the [facility] could not itself be considered `waters of the United States"' and that this would affect whether Plaintiffs could also prevail under Healdsburg. Hawaii 1, 24 F. Supp. 3d at 1001. But the court erred in attempting to apply Healdsburg because the jurisdictional status of groundwater itself is irrelevant to whether discharges that move through groundwater to jurisdictional waters require NPDES permits. P& Case: 15-17447, 05/31/2016, ID: 9997388, DktEntry: 40, Page 33 of 45 II. THE COUNTY Is LIABLE FOR UNPERMITTED DISCHARGES DUE TO THE "DIRECT HYDROLOGICAL CONNECTION" BETWEEN THE GROUNDWATER AND THE OCEAN. Discharges of pollutants from a point source that move through groundwater are subject to CWA permitting requirements if there is a direct hydrological connection between the groundwater and a jurisdictional surface waters Ascertaining whether there is a direct hydrological connection is a fact -specific determination. 66 Fed. Reg. at 3017. To qualify as "direct," a pollutant must be able to proceed from the point of injection to the surface water without significant interruption. Relevant evidence includes the time it takes for a pollutant to move to surface waters, the distance it travels, and its traceability to the point source. These factors will be affected by the type of pollutant, geology, direction of groundwater flow, and evidence that the pollutant can or does reach jurisdictional surface waters. Id. Here, the district court correctly held that the County discharged pollutants to the ocean through groundwater. In Hawaii I, the court 6 Some courts refer to a "hydrological connection." The more accurate formulation, however, is a "direct hydrological connection," which recognizes that some connections are too circuitous and attenuated to be under the CWXs purview. of Case: 15-17447, 05/31/2016, ID: 9997388, DktEntry: 40, Page 34 of 45 determined that a direct hydrological connection exists between the groundwater and the ocean. The tracer -dye study clearly established that the discharges moved from wells 3 and 4 to the ocean in relatively short order.? Hawaii I, 24 F. Supp. 3d at 984. The study concluded that after 84 days, the dye began to appear along the North Kaanapali Beach, half a mile from the facility. Id. The tracer -dye study also estimated that 64% of the treated effluent from wells 3 and 4 followed this route to the ocean. Id. Although the court's ultimate conclusion was correct, the court's alternative explanation for the County's liability under the "significant nexus" standard from Rapanos and Healdsburg was erroneous. Hawaii I, 24 F. Supp. 3d at 1004. Rapanos and Healdsburg applied the "significant nexus" standard in determining whether the receiving waters were "waters of the United States." In contrast, here, there is no dispute that the Pacific Ocean (the receiving water in this case), as a "territorial sea," is a "navigable water" under the CWA. This Court 7 Although this tracer -dye study simplified the analysis, such studies are not the only means of demonstrating a direct hydrological connection. It also is not necessary to trace the exact pathway that the pollutants take to establish that a direct hydrological connection exists. 27 Case: 15-17447, 05/31/2016, ID: 9997388, DktEntry: 40, Page 35 of 45 should clarify that the "significant nexus" standard has no relevance here. In Hawaii II, the district court correctly held the County discharged pollutants from wells 1 and 2 to the ocean through groundwater. But the court's opinion did not go into great detail about the movement through groundwater because the County "expressly conced[ed] that pollutants introduced by the County into wells 1 and 2 were making their way to the ocean" and "acknowledge [d] that there is a hydrogeologic connection between wells 1 and 2 and the ocean." Hawaii II, 2015 WL 328227, at *2. There was additional evidence that a direct hydrological connection existed between wells 1 and 2 and the Pacific Ocean. First, the tracer -dye study models indicated that in some circumstances treated effluent from well 2 would move along flowpaths that were similar to those traveled by the dye injected into wells 3 and 4 and would emerge at the same submarine springs. SER 237, 240, 243. Because wells 3 and 4 are located between the springs and well 2, the flowpath for these discharges would be affected by the amount of effluent injected into each well. SER 237. When most of the effluent was PIT -2 Case: 15-17447, 05/31/2016, ID- 9997388, DktEntry: 40, Page 36 of 45 injected into wells 3 and 4, the effluent from well 2 would travel northwesterly from the wells and not toward the springs; however, when well 2 received all of the effluent, the study indicated that the discharges would emerge at the springs. SER 240, 243. There was no dispute that given the proximity of wells 1 and 2, the modeling for well 2 predicts the pathways for discharges from well 1. ER 443, SER 189. Second, Plaintiffs' expert stated that the effluent discharged from wells 1 and 2 "will be conveyed ... relatively quickly (i.e., with first arrival at the ocean in a matter of months)" and concluded that "[s]ince the aquifer material and hydraulic gradient in the area of all four ... wells are similar, the groundwater flow will also be similar." SER 183. Although the County's expert argued that the point of entry for pollutants into the ocean from wells 1 and 2 could not be identified, the County did not dispute that the study showed effluent emerging at the same springs where the effluerit from wells 3 and 4 emerged. Haw. Wildlife Fund v. Cty. of Maui, No. 12-198, ECF No. 136, at 16. Any fears about the implications of point -source discharges to jurisdictional surface waters through groundwater with a direct hydrological connection being subject to NPDES-permit requirements 29 Case: 15-17447, 05/31/2016, ID: 9997388, DktEntry: 40, Page 37 of 45 are unwarranted. Op. Br. at 43-44. EPA and states have been issuing permits for this type of discharge from a number of industries, including chemical plants, concentrated animal feeding operations, mines, and oil and gas waste -treatment facilities. See, e.g., NPDES Permit No. NM0022306, available at https://www.env.nm.gov/swqb/Permits/; NPDES Permit No. WA0023434, available at http s://yosemite . ep a. gov/r 10/water. nsf/NPD ES+Permits/Curre ntOR&W A821. Further, only those discharges that move through groundwater with a direct hydrological connection to surface waters are affected. That not all discharges through groundwater are subject to NPDES- permit requirements is shown by cases where the hydrological connections were too attenuated. In McClellan Ecological Seepage Situation (MESS) v. Weinberger, the court agreed with the plaintiff that discharges through groundwater may be subject to the CWA and allowed the parties to submit evidence on the issue. 707 F. Supp. 1182, 1196 (E.D. Cal. 1988). Based on evidence indicating that it would take "literally dozens, and perhaps hundreds, of years for any pollutants in the groundwater to reach surface waters," the court found that there Ell Case: 15-17447, 05/31/2016, ID, 9997388, DktEntry: 40, Page 38 of 45 were no regulated discharges. MESS v. Cheney, 763 F. Supp. 431, 437 (E.D. Cal. 1989). And even after allowing the plaintiff an opportunity to provide more testimony at trial, the court ruled that the plaintiff had failed to meet its burden. MESS v. Cheney, No. 86-475, 20 Envtl. L. Rep. 20,877 (E.D. Cal. Apr. 30; 1990), vacated on other grounds, 47 F.3d 325, 331 (9th Cir. 1995). Likewise, in Greater Yellowstone Coalition v. Larson, evidence indicated that the connection to surface waters was too attenuated. 641 F. Supp. 2d 1120 (D. Idaho 2009), aff'd 628 F.3d 1143, 1153 (9th Cir. 2010). In that case, federal agencies determined that a CWA Section 401 certification was not required for a mining operation. Under Section 401, "[a]ny applicant for a Federal license or permit to conduct any activity ... which may result in any discharge into the navigable waters, shall provide the licensing or permitting agency a certification from the State ... that any such discharge will comply with the applicable provisions." 33 U.S.C. § 1341(a)(1). The agencies based their determination on evidence that before reaching surface waters, the pollutants would pass through hundreds of feet of overburden and bedrock, and then travel underground through soil and rock for one to 31 Case. 15-17447, 05/31/2016, ID 9997388, DktEntry 40, Page 39 of 45 four miles. Greater Yellowstone, 641 F. Supp. 2d at 1139. Modeling predicted that the movement of peak concentrations would take between 60 and 420 years. Id. The court weighed competing evidence from the plaintiff and ultimately deferred to the agencies' determination that the hydrological connection was too attenuated. Id. at 1141. Unlike MESS and Greater Yellowstone, in which the connection was too attenuated, the discharges here resulted from a direct hydrological connection, and thus require a permit. III. THE DISTRICT COURT CORRECTLY HELD THAT THE COUNTY HAD FAIR NOTICE FOR PURPOSES OF CIVIL PENALTIES. In the Argument section of its brief, the County maintains that this Court should direct the district court to set aside any civil penalties "imposed on the County regardless of the outcome of the challenge to the district court's liability rulings" because it lacked fair notice. Op. Br. at 47. As an initial matter, the County would seemingly be precluded from appealing the fair -notice issue as to civil penalties because it stipulated to their amount in the settlement agreement. To the extent that the County has reserved its right to appeal the issue, however, the County's argument lacks merit. 32 Case: 15-17447, 05/31/2016, ID: 9997388, DktEntry: 40, Page 40 of 45 This Court has held that a party may not be deprived of property through civil penalties without fair notice. See United States v. Approximately 64,695 Pounds of Shark Fins, 520 F.3d 976, 980 (9th Cir. 2008). To provide notice, "a statute or regulation must `give the person of ordinary intelligence a reasonable opportunity to know what is prohibited so that he may act accordingly."' Id. This Court looks first to the language of the statute when determining whether a party had fair notice. Id. As discussed above, Congress used broad language in the CWA in defining the discharge of pollutants, and that expansiveness provides a reasonable opportunity for a person to know what the statute prohibits. The breadth of that language is only bolstered by the intent of the CWA. Moreover, EPA has made multiple public statements in rulemaking preambles that consistently described its interpretation that discharges of pollutants to jurisdictional surface waters through groundwater with a direct hydrological connection are subject to NPDES permitting under the CWA. Further, with respect to specific communications with the County, EPA sent two letters to the County in early 2010. In January 2010, EPA stated that it was "investigating the 33 Case: 15-17447, 05/31/2016, ID: 9997388, DktEntry, 40, Page 41 of 45 possible discharge of pollutants to the coastal waters of the Pacific Ocean along the Kaanapali coast of Maui." SER 5. This investigation was spurred in part by a 2007 study concluding that much of the nitrogen in Kaanapali coastal waters came from the County's facility and a 2009 study that found the same nitrogen signature and other "wastewater presence" in the ocean. Hawaii 1, 24 F. Supp. 3d at 984. The letter continued: "In order to assess the impact of the [facility's] effluent on the coastal waters and determine compliance with the Act, EPA is requiring the County to sample the injected effluent, sample the coastal seeps, conduct an introduced tracer study, and submit reports on findings to EPA." SER 5. EPA required this sampling, monitoring, and reporting pursuant to CWA Section 308, under which "the [EPA] Administrator shall require the owner or operator of any point source" to provide the information. 33 U.S.C. § 1318(a)(A). The letter provided notice that there was evidence suggesting a hydrological connection. In March 2010, EPA responded to the County's request for a UIC permit renewal under the SDWA "by informing the County that recent studies `strongly suggest that effluent from the facility's injection wells is discharging into the near shore coastal zone of the Pacific Ocean." 34 Case: 15-17447, 05/31/2016, ID. 9997388, DktEntry. 40, Page 42 of 45 Hawaii I, 24 F. Supp. 3d at 984 (quoting ER 122). As a result, EPA required the County to apply for a CWA Section 401 water -quality certification for its injection facilities as a prerequisite to EPA's issuance of a new UIC permit. ER 121-22; see 33 U.S.C. § 1341(a). The County's assertion that this letter did not put it on notice of potential CWA liability because the certification was related to its UIC permit rather than any obligations under the NPDES program is unavailing. Op. Br. at 56-57. A UIC permit does not preclude the need for a NPDES permit where required, and the March 2010 communication reiterated EPA's position that the discharges might be covered by the CWA, depending on the results of the ordered sampling, monitoring, and reporting. The County was on fair notice. In any event, fair notice is only one of many factors informing a civil -penalty amount, see 33 U.S.C. § 1319(d), and thus the County's argument that the penalty should be set aside for lack of fair notice alone is flawed. 35 Case: 15-17447, 05/31/2016, ID. 9997388, DktEntry: 40, Page 43 of 45 CONCLUSION For the foregoing reasons, the district court's judgment should be affirmed. OF COUNSEL: KARYN WENDELOWSKI U.S. Environmental Protection Agency Office of General Counsel Washington, D.C. May 31, 2016 90-12-14672 Respectfully submitted, JOHN C. CRUDEN Assistant Attorney General /s / Frederick H. Turner FREDERICK H. TURNER AARON P. AVILA R. JUSTIN SMITH Attorneys, U.S. Dep't of Justice Env't & Natural Resources Div. P.O. Box 7415 Washington, DC 20044 (202) 305-0641 frederick.turner@usdoj.gov 36 Case: 15-17447, 05/31/2016, ID: 9997388, DktEntry: 40, Page 44 of 45 CERTIFICATE OF COMPLIANCE WITH TYPE -VOLUME LIMITATION, TYPEFACE REQUIREMENTS, AND TYPE -STYLE REQUIREMENTS This brief complies with the type -volume limitation of Fed. R. App. P. 32(a)(7)(B) (for amicus briefs as provided by Fed. R. App. P. 29(d)) because it contains 6,904 words, excluding the parts of the brief exempted by Fed. R. App. P. 32(a)(7)(B)(iii). This brief complies with the typeface requirements of Fed. R. App. P. 32(a)(5) and the type -style requirements of Fed. R. App. P. 32(a)(6) because it has been prepared in a proportionally spaced typeface using Microsoft Word 14 -point Century Schoolbook. 37 /s / Frederick H. Turner FREDERICK H. TURNER U.S. Department of Justice Env't & Natural Resources Div. P.O. Box 7415 Washington, DC 20044 (202) 305-0641 frederick.turner@usdoj.gov Case: 15-17447, 05/31/2016, ID: 9997388, DktEntry: 40, Page 45 of 45 CERTIFICATE OF SERVICE I hereby certify that on May 31, 2016, I electronically filed the foregoing brief with the Clerk of the Court for the United States Court of Appeals for the Ninth Circuit using the appellate CM/ECF system, which will serve the brief on the other participants in this case. I s /Frederick H. Turner FREDERICK H. TURNER U.S. Department of Justice Env't & Natural Resources Div. P.O. Box 7415 Washington, DC 20044 (202) 305-0641 frederick.turner@usdoj.gov Wo