Loading...
HomeMy WebLinkAboutNCD122263825_20150827_JFD Electronics - Channel Master_FRBCERCLA SPD_Draft Proposed Buffer Zone Around Area A for LUR-OCR-~---.--• •· . . WArERS EDGE ENVIRONMENTAL, LLC . , 4901 WATERS EDGE DRIVE, SUITE 201 • RALEIGH, NC 27606 • PHONE 919.859.9987 • FAX919.859.9930 Date To From Subject , MEMORANDUM l\'b';)\0; lL. -~\ August 27, 2015 :f?f!>_,.-(J.C"'i'.,\¢'0 ' "' \'.>-' ,-,r;:,\S '.". st? '-Ms. Mary Siedlecki-NCDENR -" \\S ; \ .. ~oo . \\ . \\"'L <;,ec\\O . / Mr. Phil Rahn-Waters Edge {b~ ,~-.s\C /, t\2~. _,,,:J/ . l..{(,.r, -~·,-: l _,., Draft Proposed Buffer Zone Around Area A for LUR Submittal~· Daimler Facility Cleveland, North Carolina I have been tl)'ing to research a method of calculating a "Buffer Zone" around Area A for use as an Area B for the land use restrictions (LUR) at the Daimler facility in Cleveland, North Carolina You had referred me to another facility (Schneider Electric site in Knightdale, NC) which had undergone a similar simulation as an example of something that would be acceptable to NCDENR. We were able to download the Schneider Facility Remedial Action Plan supplement (Appendix F), and did determine that the consultant had undergone integrating a MODFLOW calibrated groundwater model which we believe was primarily used to predict the downgradient extent of contaminant impact but then was also used for this capture zone simulation. The corresponding difficulty is that this MODFLOW calibrated groundwater model is a considerable costly undertaking and we again believe that the primary use of this model was for contaminant plume extent simulations. Our belief is that there may be other less expensive methods of calculating this capture ~= zone that would not involve a calibrated groundwater model. Based on some additional research, we did find an article titled Domestic Well Capture Zone and InJluence of_Jf)f, •• Gravel Pack Length (see Apriendix A). On Page 3, Paragrarih 2, it states "where regional grounClwat#flowisdbminant and-local rechargeTnegligib'te,_the cariture zone ofa domestic well can also be easily comriuted if the well fulliJ:>eootralae 119ui"fer system_ ordoes not stronglY. affect regional groundwater flow. The widih, w, ofihe capture zone .... =,=.;,~ ' -, of a domestic well is then obtained by simple mass balance (Todd 1980): w = Ql(I'*i) where Q is the pumping rate, T is the aquifer transmissivity, ?P.d~Liscthe::regional::::i • V hydraulic gradient. Pumping Rate (0)-We did find a reference for the pumping rate (yield) in an article titled Statistical Analysis Relating Well Yield to Construction Practices and Siting of Wells in Domestic Well Capture Zone Daimler Facility Cleveland, North Carolina August 27, 2015 Page 2 of2 the Piedmont and Blue Ridge Provinces of North Carolina (see Appendix B). This article cross referenced well yields to both hydrogeologic conditions and topography and we were able to find that the average well yield for the hydrogeologic conditions (only past well log that referenced stratigraphy was from MW-I (boring log included as Appendix C) which described the bedrock type as a meta-diorite [Mil]-see Appendix B) and topography (we believe that this site would be described as being situated on a hill or ridge) identified at the Daimler site was an average of 13.3 gallons per minutes (GPM- Appendix 8-Page A23-Table 7). Converting 13.3 GPM to cubic meters per minute is as follows: 13.3 GPM * 0.003785 m3/gallon = 0.05034 m3/minute =72.49 m3/day =26,458.97 m3/year Transmissivity m-We found a reference titled Classification of Transmissivity Magnitude and Variation (Appendix D-Page 231 Table 1-Classification of Transmissivity Magnitude) which provides an "intermediate" coefficient of transmissivity at I 00 m2/day. Converting I 00 m2/day to m2/year is as follows: 100 m2/day x 365 days/year= 36,500 m2/year Hydraulic Gradient (i). The most recently calculated hydraulic gradient from the May 20 I 5 groundwater report is 0. 022 f1/ft. Using these variables, the derived formula would be as follows: w = 26,458.97 m3/year/(36,500 m2/year * 0.022) = 32.95 m * 3.28 feet/meter = 108.08-foot capture zone width We have depicted a 110' capture zone width for Area Bon Figure 1 for review purposes. We would like to use this as our Area B Zone for the LUR submittal and would ask that you review this submittal to see if this is acceptable. FIGURES APPENDIX A Domestic Well Capture Zone an :, Influence of the Gravel Pack Length 2 Judith E. Hom and Thomas Harter' 3 Department of Land, Air, and Water Resources 4 University of California 5 Davis, CA 95616-8629 6 • (corresponding author; ThHarter@ucdavis.edu; 530-752-2709) 7 8 Accepted for publication in the journal "Ground Water", September 2008 9 10 Abstract 11 Domestic wells in North America and elsewhere are typically constructed at relatively shallow 12 depths and with the sand or gravel pack extending far above the intake screen of the well (shallow 13 well seal). The source areas of these domestic wells and the effect of an extended gravel pack on the 14 source area are typically unknown and few resources exist for estimating these. In this paper, we use 15 detailed, high-resolution groundwater modeling to estimate the capture zone (source area) of a 16 typical domestic well located in an alluvial aquifer. Results for a wide range of aquifer and gravel 17 pack hydraulic conductivities are compared to a simple analytical model. Correction factors for the 18 analytical model are computed based on statistical regression of the numerical results against the 19 analytical model. This tool can be applied to estimate the source area of a domestic well for a wide 20 range of conditions. We show that an extended gravel pack above the well screen maY. co11trib_ute 21 significantly to the overall inflow to a domestic well, especially in less permeable aquifers, where 22 that contribution may range from 20% to 50%; and that an extended gravel pack may lead to a 23 significantly elongated capture zone, in some instances nearly doubling the length of the capture I Hom 24 zone. Extending the gravel pack much above the intake screen therefore significantly increases the 25 vulnerability of the water source. 26 27 Introduction 28 Most households in rural areas of the United States, outside the service area of incorporated cities, 29 rely on domestic wells for their water supply (McCray 2005, U.S. EPA 1997). And many of these 30 domestic wells are constructed with a well-screen at depth and a sand or gravel pack that extends 31 upward to the mandatory minimum depth of the well seal, which is dictated by local and state 32 guidelines. A question commonly asked by homeowners is: Where does our water come from? The 33 capture zones (also referred to as the source area or recharge area) of domestic wells are rarely 34 determined. Attention has instead focused on public supply wells and their capture zones as these are 35 regulated through U.S. EPA's source water protection program. Domestic wells, typically serving a 36 single family, are often constructed to relatively shallow depths when compared to public or 37 municipal water supply wells (Burow et al. 2004). 38 39 40 41 42 43 44 45 Methods for delineating well capture zones range from very simple to very complex. In general, the various approaches fall into four categories (Harter, 2008): 1.c1aeometric or graphical methods involve the use of a pre-determined fixed radius without any special consideration of the flow system, or the use of simplified shapes_ that have been pre- calculated for a range of pumping and aquifer conditions. 2. ~alytical methods allow calculation of distances for protection zones using equations that can be solved using a hand calculator or microcomputer spreadsheet program. 2 Hom 46 3. Hydrogeologic mapping involves identifying the recharge zone and the source zone based on 47 geomorphic, geologic, hydrologic, and hydrochemical characteristics of an aquifer. 48 4. Computer modeling methods involve devising, calibrating, and applying complex analytical or 49 numerical models that simulate groundwater flow and contaminant transport processes. 50 The long-term average pumping rate of domestic wells typically ranges from less than 4 L/min [ I ===== 51 gallon/min) to 20 L/min (5 gallon/min]. Using the graphical method employed by California's 52 Drinking Water Source Assessment and Protection (DWSAP) Program (California OHS, 1999), for 53 example, the default source area of a domestic well pumping 1,233.5 m3/year (I acre-foot per year, 54 the typical annual consumption ofa U.S. single family household) is a circle with a radius of 15 m 55 (~50 ft) for areal recharge of 450 mm/year (typical for very humid areas or rural residences in semi-<:::e:=:...,. 56 arid areas surrounded by irrigated lawn and fields) or with a radius of 31 m ( ~ 100ft) at a recharge 57 rate of 100 mm/year (typical of many semi-arid regions). This simple geometric approach neglects 58 the effects of the regional groundwater flow on the capture zone of a domestic well. 59 60 On the other hand, where regional groundwater flow is dominant and local recharge is negligible, the 61 c~(Jture zone of a domestic well can also be easily computed if the well fully penetrates the aquifer 62 sxstem or does not strong!y affect regional groundwater flow. The width, w, of the capture zone of a 63 domestic well is then obtained by simple mass balance (Todd 1980): 64 w = QI (T* i) 65 where Q is the pumping rate, Tis the aquifer transmissivity, and i is the regional hydraulic gradient. 66 For example, at a relatively low transmissivity, T, of 10 m2/d, a regional groundwater gradient of 67 0.5% and a pumping rate of 1,233 m3/year, the width of the capture zone is approximately 60 m 3 Horn 68 (~200 feel). Al values of Ttypical for productive aquifers, the width of the capture zone is often on 69 the order of I m -IO m ( ~ 3 feet -~ 30 feet) or even less. 70 71 Both, the geometric approach and equation (I) above provide simple approximations for extremely 72 idealized conditions. Here, our objective is lo determine the capture zone of a domestic well with a 73 sand or gravel pack, completed in an unconfined aquifer,1where both, recharge and regional 74 groundwater flow are significant. We use high-resolution computer simulations to determine the 75 source area and to explicitly determine the influence of the gravel pack on the well capture zone. For 76 reference, we compare those to a simple analytical model of the capture zone for a low-producing 77 well in an unconfined aquifer with recharge. Our study's focus is on rural domestic wells in irrigated 78 agricultural regions, e.g., of the Southwestern United States, where significant recharge is due to 79 irrigation return flows and much of the groundwater production is for irrigation purposes. Our 80 findings have general implications that are independent of this particular climate scenario. 81 82 Conceptual Framework 83 Domestic wells in rural areas are assumed to be completed near the uppermost portion of a regional 84 aquifer system. Furthermore, we assume that Dsignificant downward gradientfexists in the regional 85 aquifer system due to recharge at the water table and due to significant groundwater production 86 (mostly for irrigation) from the deeper portions of the aquifer system (e.g., Belitz and Phillips, 87 1995). Burow et al. (2004), for example, report typical recharge rates in irrigated areas in the San 88 Joaquin Valley, California, lo be on the order of 550 -750 mm/a with the majority of recharge 89 originating from irrigation return flows. For simplicity, regional groundwater flow is considered lo 90 be uniform around the source area of the domestic well and at steady-state. The superposition of 91 regional groundwater flow with the downward gradient induced by water table recharge and deeper 4 Hom 92 gro1IDdwater production yields a gro1IDdwater flow field that is vertically inclined relative to the 93 slope of the water table (Figure 1). 94 95 96 97 98 99 100 IOI 102 103 104 105 106 107 108 109 110 Ill plan view cross soction .-capture zone-1 width o/\ . · · \ 1--lenglh ' --< I ! I x. C Cl> e u UI a) z, b) L--'--'--...V t--whole capture zone ----t ! elongated part ! main part ! Figure 1: Conceptual framework of gro1IDdwater flow towards a partially penetrating domestic well (a) without and (b) with a gravel pack that extends for several meters to several tens of meters above the well screen. Top: plan view, bottom: cross-sectional view. Regional gro1IDdwater flow is from right to left with a vertical flow component controlled by IIDiform recharge at the top and aquifer pumping from large production wells dispersed in the deep part of the aquifer below. The aquifer bottom is assumed to be much deeper than the typical depth of the (relatively shallow) domestic well; I: length, w: width. All other symbols: see text for details. A simple method to compute the approximate source area size and location is available, if we neglect the effects of the gravel pack and the effects of domestic well pumping, Q, on the local gro1IDdwater flow field. Then, the source area location is obtained from the length and depth of the domestic well screen, and from the angle, w, of gro1IDdwater flow relative to the slope of the water table (Figure la): Zo-Zl X1 = X w + --(3) tanw ,,,.. = x, -x. (4) 5 Horn I 12 113 114 115 116 117 118 119 A,,_0 = QI R (5) A w = --1!!!!!.. (6) ,1-0 I ,,.. where Xw is the location of the well (along the regional groundwater gradient), Xh is the location of the downgradient edge of the recharge (source) area, x, is the location of the updgractient edge of the recharge area, z0 is the elevation of the water table, Zh is the elevation of the top the well screen, z, is the elevation of the bottom of the well screen (Figure la), l,1,eo, w,heo, A,heo are the theoretical length, width, and area of the recharge zone, and: tan w =RI (Kh • i) (7) 120 where R is the uniform recharge rate, Kh is aquifer hydraulic conductivity, and i is the regional 121 hydraulic gradient. Equations (2)-(7) provide a simple analytical model to determine the capture 122 zone of a domestic well in an unconfined aquifer with uniform flow, recharge, and deep production. 123 124 To account for the influence of domestic well pumping on the local groundwater flow system around 125 the well and to account specifically for the influence of the gravel pack on the recharge area (Figure 126 lb), we constructed a numerical model, described in the next section. 127 128 Modeling Methods 129 The capture zone of a domestic well with a gravel pack is computed for a fully three-dimensional 130 steady-state groundwater flow field. The steady-state head and flux distribution are computed using 131 the MOD FLOW groundwater flow model (McDonald and Harbaugh 1988). The capture zone 132 corresponding to a particular groundwater flow solution is delineated using the backward particle 133 tracking model MODPATH (Pollock 1994). 134 6 Hom 135 Briefly, MODFLOW solves the steady-state groundwater flow equation 136 VK"vh=O (8) 137 where his the hydraulic head, by using a fully three-dimensional block-centered finite difference 138 scheme for the user-specified boundary conditions, K is the hydraulic conductivity tensor. In the 139 following simulations pumping induces only a small drawdown of the piezometric surface, so the 140 linear flow model (8) is sufficiently accurate for our purposes. We effectively invoke the Dupuit 141 assumption equivalent to the MODFLOW "unconfined layer" algorithm. There, the unconfined layer 142 thicknesses are set constant and only updated iteratively. From the hydraulic head solution, 143 MODFLOW also computes the flux, q, across each of the six faces of each finite difference cell in 144 the modeling domain. The flux solution becomes input to MODPATI-1, which computes backward 145 particle travel paths given the linear groundwater velocity, v = qln, where n is the effective porosity, 146 across each finite difference cell face. Starting locations for backward particle paths are user-defined. 147 MODPA TI-I uses a semi-analytical linear interpolation scheme to compute a spatially continuous 148 particle path (Pollock 1994). 7 Hom 149 150 151 152 153 154 155 156 157 158 159 160 161 a . I ~ I .. I "C C: :::, I 0 .rJ I "C I .. ., I .c . ... I C: B E '" I 91 C: I 0 u ' ' ' I II I I -~ b) ----x-direclion T E i l 1~ I -~ I 1 I I I ' 387m ~· .. "C . C: :::, . 0 · . .c ."C ' ii .c C·: B, ,n . C: 8 \ I _ _/ wall (888 Figura 4) conatant flux boundaiy - Figure 2: Model grid in (a) cross-sectional view aty = 0 (Vertical exaggeration= 4.2x) and (b) in plan view. Due to the symmetiy of the flow field, the model domain simulates only half of a well and half of the capture zone. The well and gravel pack are very finely discretized. A close-up view of the model around the well screen is shown in Figure 4. Our modeling domain is a finite difference grid with 141,750 cells of which 137,937 are active. The modeling domain is 58 m high, 387.23 m long and 59.695 m wide and consists of 45 rows, 90 columns, and 35 layers. The modeling domain takes advantage of the symmetry in the well flow field, which is symmetric across the x-axis (mean flow direction) centered on the domestic well (y = 0, see below). The model is therefore designed to model only one-half of the well capture zone (Figure 2). The second half of the well-capture zone mirrors the first half Grid spacing is non- 8 Hom 162 uniform in both the vertical and horizontal direction. Vertical grid spacing varies from I mat the 163 elevation of the well screen to 4 m elsewhere (Figures 2, 3). Horizontal grid-spacing varies from 164 0.0 I m near the well and in the gravel pack to nearly 20 m near the model boundaries. The 165 horizontal increase in cell-size between adjacent rows or columns of the finite-difference grid is set 166 to not exceed 50 % of its width. 167 168 The hydraulic gradient along the x-axis is produced by defining a constant head boundary of 58.00 169 m to the exterior block of cells al the upgradienl vertical side of the model and a constant head of 170 57.61 mat the downgradient vertical side of the model (Figure 2). This is equivalent to a hydraulic I 71 gradient of 0.0018, which is typical for the study area. The other two vertical planes of the model are 172 assigned no-flow boundary condition: the vertical plane adjacent to the well half is a symmetry 173 plane. The vertical plane opposite of the half well is at sufficient distance to the well that the local 174 effect of pumping on the groundwater flow field can be neglected and flow is parallel to regional 175 groundwater flow. The average (steady-state) recharge rate is set to 0.669 m/year, a value typical for 176 semi-arid, irrigated agricultural regions such as the Modesto Area, San Joaquin Valley, California. 177 The bottom of the model domain is considered permeable and open to the regional aquifer system 178 below. It is assigned a uniform constant (downward) flux boundary condition, with total outflow 179 across the bottom boundary set equal to the difference between the total recharge inflow at the top 180 and the well outflow rate. In this way we implicitly enhance our model to greater aquifer depths. In 181 the Modesto Area large irrigation wells up lo a depth of almost 370 m below land surface pump 182 large amounts of water and produce a vertical flow component, even through a confining clay unit 183 above the irrigation wells (Burow el al. 2004). 184 9 Hom l 85 The well construction was chosen to be representative of domestic well construction in the San l 86 Joaquin Valley, California (e.g., Burow et al., 2004). The model well has a total depth of 56 m below 187 the water table. A seal to 18 m below the water table overlies a 30 m long gravel pack around a l 88 blank well casing. The casing has a diameter of 0.2 m. The perforated well screen is located at 48 m 189 to 55 m below the water table, followed by a conceptual well sump from 55 m -56 m. Casing and 190 screen are surrounded by a 0.09 m thick gravel pack. The total borehole diameter is 0.38 m. Due to 191 the relatively low pumping rate, the well-loss and skin effect are assumed to be negligible. Inflow 192 along the screen is computed by the model and non-uniformly distributed. 193 194 The grouted well seal above the gravel pack and the well casing are modeled as "no-flow" cells 195 (black cells in Figure 3). The pump is simulated by 74 "well" cells inside the casing. They are 196 located significantly above the top of the screen, opposite of the well seal bottom, which creates an l 97 upward flow inside the screen and casing. The MODFLOW "well" package is used to simulate the 198 pump cells (light-grey cells in Figure 3). The total pumping rate of the domestic well is 3.5 m3/d, I 99 half of which is uniformly distributed across the individual "well" cells at the top of the casing. 200 Flow inside the model well casing was modeled by approximating the flow with eq. (8) using very 201 high hydraulic conductivity. The gravel pack (grey cells in Figure 3) is modeled by choosing a 202 separate hydraulic conductivity that is higher than that of the surrounding aquifer and ranges 203 between 50 and 1000 [mid] (Table 1). Modeling the pump inside the well allows the model to 204 properly distribute the flow across the well screen, with screen inflow highest near the top of the 205 screen and lowest at the bottom of the screen. Hom E 18 •, IDIDIIIIDlfflfflllllllfflllll 1111111 · H · IIHIIIIIIIIIIIII WIii] ~ .0 calls ••••• • ■ ••• • • • ■ •••• . . . • • • • • • ii •• • • • • gravel pack . I ■ ■ I I ■ I I I I ■ ■ I I I I I I I I I I I I I I I I I I • •••••••••••••••••••• • •••• ■ •••••• ■ • • • • • ■ ■ I I ■ I I I ■ I I I I I I I ■ ■ I I I I I ■ ■ ■ ■ ■ I I ■ I I I I ■ I I I I I I I ■ I ■ I I I ■ I I I .......... ■ I ■ ■ I ..... 1111-rth-~g pcilnt of parlk:lall ......... ■ ■ ■ ■ I I I ■ I ■ ■ I I I I ■ I I I I I I I I I I I I I I I I ■ I I I I I I ■ I I I I I I I I I . . . . . . . I I I I I ■ I ■ I ■ I I I I I I I I I I I ■ ■ I I I I ■ I ■ ■ ■ ■ I ■ ■ I I I ■ I I I ■ I ■ I I I I I I I I I I Ill I I ■ ■ I I I I I I I I I I ■ I I I I ■ ■ I I ■ ■ I I I I ■ I ■ I I I I I I ■ ■ I ■ ■ I I I I ■ I I I I ■ I I I I I I I I I I ■ ■ I ■ I I ■ I I I I ■ I I I I I I I ■ I ■ I I I ■ ■ ■ ■ I I . . .................... . ■ I I I ■ I ■ ■ I I I I I I ■ I I I II I I I I I ■ I ■ I I I ■ I I I I I I ■ ■ I I I I I I I I I ■ ■ ■ I ■ I I I I I I I I I I ■ I 206 . 0.38 m 1-----0.38 m -------l 207 Figure 3: Model well configuration and grid discretization around the well. Left: Cross-section at 208 the model boundary (y = 0). Right: Plan view at the land surface (right top), at the top layer of the 209 casing containing the well cells (right center), and at the screen elevation (right bottom). Black cells: 210 casing and well seal (impermeable). Grey cells: gravel pack. Dark grey cells: constant flux boundary 211 cells at the model bottom. Grey dots in the lower left panel indicate the starting location for 212 backward particle tracking. 213 214 11 Hom 215 The hydraulic conductivity, Kh, is assumed to be isotropic in the horizontal plane, while the vertical 216 aquifer hydraulic conductivity, K.., is lower, as typically observed in alluvial aquifers (e.g., Phillips 217 et al., 2007). Two representative anisotropy ratios, Kh/K, = 5 and 2, were chosen to bracket a 218 representative range typically found in alluvial aquifers (ibid.). The gravel pack itself is assumed to 219 have a completely isotropic hydraulic conductivity, Kg, that is larger than Kh, For illustration and 220 application purposes, we modeled well capture zones for a wide range ofrepresentative values for 221 the horizontal hydraulic conductivity, Kh, and the gravel pack hydraulic conductivity, Kg, and for two 222 anisotropy ratios (Table 1 ). Kh K, K I 0.2 50,125,250,500,750, 1000 I 0.5 50,125,250,500,750 1000 3 0.6 50 125 250 500 750 1000 3· 1.5 50 125 250 500 750 1000 5 I 50 125 250 500 750 1000 5 2.5 so, 125,250,500,750, 1000 10 2 so, 125,250,500,750, 1000 10 5 50 125 250 500 750 1000 30 6 50 125 250 500 750 1000 30 15 50 125 250 500 750 1000 100 20 125 250 500 750 1000 100 50 125,250,500,750, 1000 300 60 500,750 1000 300 150 500 750, 1000 223 224 Table 1: Model configurations with various 225 combinations of the horizontal hydraulic 226 conductivity, Kh, the vertical hydraulic 227 conductivity, K,, and the gravel pack hydraulic 228 conductivity, Kg-All values are in units of 229 [mid]. 230 231 Results 232 Head contour configurations in the aquifer around the domestic well are highly dependent on the 233 aquifer and gravel pack hydraulic conductivities. Cross-sectional head contour lines along the 12 Hom 234 regional flowpath are vertical under strictly_regional flow with no recharge and no pumping. As 235 expected from the analytical model above, the modeled contours deviate from the vertical due to the 236 vertical flow component imposed by the recharge at the top of the model area and the regional 237 pumping below the modeled zone. Contour lines increasingly deviate from the vertical alignment 238 with smaller and smaller ratios of Kh IR (Figure 4). In addition, in aquifers with relatively low 239 hydraulic conductivity, the domestic well creates a distinct zone oflocal influence in the aquifer 240 around the well screen, whereas the influence is minimal in the highly permeable aquifer. The 241 anisotropy of the aquifer hydraulic conductivity creates significant flow zonation: much of the 242 impact of domestic well pumping on the pressure field is seen at the elevation of the well screens, 243 especially for those cases with the higher aquifer anisotropy. Another distinct horizontal zone is 244 created by the top of the gravel pack. The higher the gravel pack hydraulic conductivity (relative to 245 Kh), and the higher the aquifer anisotropy ratio, Kh!K,, the more pronounced is the effect that the 246 transition between the top of the gravel pack and the annular seal has on the head contour lines (e.g., 247 Figure 4). Inflow to the well varies non-uniformly along the screen. It is highest near the top of the 248 screen, which is nearest to the pump intake inside the well-casing. The difference between the screen 249 inflow at the top (layer 26) and the screen inflow near the bottom (usually in layer 31 just above the 250 bottom of the layer) varies from approximately 45% for highly permeable aquifers to more than 251 100% for veiy low permeable aquifers with veiy high gravel pack K8. This is consistent with 252 analytical models (Nahrgang, 1954; Garg and Lal, 1971) and with field observations on large 253 production wells (VonHofe and Helweg, 1998). 13 Hom - 254 255 Figure 4: Head contour lines around the well for a conductivity of 10 mid, an anisotropy ratio of 2, 256 and hydraulic gravel pack conductivity of 750 mid. The heads depend on the conductivity of the 257 aquifer, the anisotropy and the relative difference in the conductivities between gravel pack and 258 aquifer. Horizontal dimension is 7 m, vertical dimension is 58 m. Due to the horizontal exaggeration 259 (12.4x) the inclination of the head contours in the regional flow field (near top of the cross-section 260 appears nearly horizontal although it is actually nearly vertical. 261 262 Corresponding to the head field, pathlines in low hydraulic conductivity aquifers are significantly 263 steeper and the capture zone is much closer to the well-head than in an aquifer with high hydraulic 264 conductivity (Figure 5). For Kh 2': 10 ml day, the modeled pathlines are in fact sufficiently flat that 265 the source area is outside the model area In those simulations, we computed the pathlines outside 266 the numerical modeling area by analytically calculating the extension of the pathlines to the water 267 table using equation (7). Also, for model scenarios with hydraulic conductivities of 1, 3, and 5 mid, 268 the pathlines in the top aquifer layer were computed from eq. (7), because MODP A lll computations 269 in the top layer were subject to numerical error. 14 Hom 270 271 The source area of the domestic well has a distinct shape composed of two features: the main 272 capture zone, a relatively large and wide oval area, corresponding to pathlines that enter the annulus 273 of the well below the top of the well-screen for horizontal delivery into the well. At the 274 downgradient (well-facing) side of this main capture zone, we observe a narrow elongated capture 275 subzone that represents those pathlines that enter the gravel-pack of the well at some distance above 276 the well screen. These pathlines capture domestic water through the high permeability field of the 277 gravel pack above the well screen (Figure lb, Figure 5). The greater the hydraulic conductivity 278 difference between gravel pack and aquifer, the higher is the relative downward flow in the upper 279 gravel pack, and the more MODPA TH virtual water particles enter the well flowing through the 280 upper gravel pack. Moreover, the steeper the particle path gradient, the higher is the highest point of 281 entry into the gravel pack of pathlines that ultimately will be captured by the well. Thus, the gravel 282 pack, where it extends to elevations much higher than the well-screen, significantly extends the 283 length of the source area towards the well, albeit within a very narrow transverse range (Figure 5). -~---== ,. " -' ., ~ ~ ~ 284 285 Figure 5: Pathlines in cross section (left) and plan view (half the well, right) with the elongated and 286 main capture zone parts for an aquifer conductivity of 10 mid, an anisotropy ratio of 2, and gravel 287 pack conductivity of 750 mid. Corresponding heads are shown in Figure 4. 15 Hom '.' 288 289 For further analysis of the capture zone location and size, we separately refer to the width and length 290 of the narrow, "elongated" part of the capture zone nearer to the well and of the "main" part of the 291 capture zone (Figure 6) from where the majority of the water originates. The simulations show that 292 the length of the elongated part increases faster than the length of the main part as horizontal aquifer 293 conductivity increases, but the gravel pack conductivity has a significant influence only on the 294 length of the elongated part (Figure 6c, d). The same is true for the width of the two capture zone 295 parts: The gravel pack conductivity has a significant influence only on the width of the elongated 296 part but little, yet discemable influence on the main part. The width of the elongated part increases 297 several-fold with gravel pack hydraulic conductivity, K8, especially in less productive (low K) 298 aquifers. By the same token, the widths of the main and elongated parts (Figure 6a, b) decrease 299 with higher aquifer conductivities (more narrow, but longer source area). For low gravel pack 300 conductivities, the width of the elongated part of the capture zone remains nearly constant, 301 regardless of aquifer conductivity 16 Hom a) 50 E 40 fil ~ 30 "i i 20 a. C -~ 10 ■ 0 3 5 10 30 100 300 c) 1200 Kh [mid] -1000 .E. II .c 800 0) C .9l 600 i::: ro 400 a. I!! C -~ 200 0 ■ • • Ii Ii -b) 12 E 10 ,5 -0 8 "i i 6 a. -0 4 .9l ro Cl 2 C .Q 0) 0 ■ • 6 0 B e 3 5 10 30 o Kg: 50 □ Kg: 125 ◊ Kg: 250 6 Kg: 500 • Kg: 750 ■ Kg: 1000 100 300 K• [mid] d) 100 ~---~-~---~ .E. 600 -g, 500 C .9l 400 i::: ~ 300 ]l 200 "' i? 100 .Q " ow:!...--m...t1.....tl------~~ 1 3 5 10 30 100 300 3 5 10 30 100 300 3m ~~ ~~ 303 Figure 6: Widths (top panels) and lengths (bottom panels) of the elongated part (right panels) and 304 the main part (left panels) of the capture zone for an anisotropy ratio ofKv: Ki,= 1 : 2. Behavior of 305 the models with an anisotropy ratio Kv : Ki, = 1 : 5 is similar. 306 307 The analytical model (eqs. 2,3) of the source area location provides good approximations of the 308 source area only in highly permeable aquifers. For aquifers with intermediate and low conductivity, 309 the gravel pack has significant influence on the distance of the downgradient edge of the capture 310 zone from the well (Figure 7a), where the source area can be as much as 90% closer to the well than 311 estimated from eq. 2. The analytical approximation of the distance from the well to the upgradient 312 edge of the source area (eq. 3) is relatively close to the numerical simulations if aquifer hydraulic 313 conductivities are above 5 mid. In those cases, the relative difference between analytical and 314 numerical model is on the order of 10% or less (Figure 7b), regardless of anisotropy ratio and gravel 315 pack hydraulic conductivity. 17 Hom a) 1,0 b) 0,9 GI el " !'l o Kg: 50 C: ! ]j 0,8 " .'9 0 = o Kg: 125 i 0,8 .!ll ◊ Kg: 250 'O 0,7 E § t. Kg: 500 .; 0 6 \'I 2 E 0,6 • Kg: 750 'I< Bl Kg: 1000 ·-' ~ 0,5 E ~ □ .. ~ 0,4 e1 0,4 ◊ Q) • 5l 0 d j 0,3 :i □ 6l ] 0,2 0 ◊ ";' 0,2 □ 0 al Ill ai 0 G ~ 0,1 I!! 'O &l 1l El ~ 0,0 ll E 0,0 3 5 10 30 100 300 3 5 10 30 100 300 316 Kh [mid] Kh [m/d] 317 Figure 7: Comparison of the distances of the source areas to the well provided by the numerical and 318 by the analytical model exemplarily for an anisotropy ratio of Kv: Ki,= 1 : 2. (a) Normalized 319 differences between the modeled and analytically calculated distances of the downgradient edges of 320 the source areas to the well. (b) Normalized differences of the distances of the upgradient edges of 321 the source areas to the well. 322 323 The simulation results show that water moves downward inside the gravel pack above the well- 324 screen from considerable distances: For Kh less than 10 mid and high gravel pack hydraulic 325 conductivities, water travels downward from as far as the top of the gravel pack, 30 m above the 326 well-screen (Figure 8). Again, the more permeable the gravel pack in the annulus, the larger the 327 above-screen capture of source water. The fraction of well pumpage that originates from capture in 328 the gravel pack above the well-screen increases as the aquifer hydraulic conductivity decreases 329 (Figure 9). In intermediate and low permeable aquifers, domestic wells with highly permeable 330 gravel packs receive from 20% to 50% of the total well flow from the extended gravel pack above 331 the screened aquifer horizon. This model result is qualitatively consistent with the field data of 332 Houben (2006), who found iron oxide incrustations in the gravel pack significantly above the top of 333 the well screen, where the incrustations were due to a significant amount of water flowing through 334 the upper part of the gravel pack. At high aquifer conductivities (Kh > 10 mid), less than 12 % of the 18 Hom 335 total domestic well flow originate from the gravel pack above the well screen. Aquifer anisotropy 336 has little influence on the height of the capture zone within the gravel pack. 337 338 The height of the gravel pack participating in flow to the well and the percent fraction of the 339 pumpage originating from the gravel pack above the screen can be expressed quantitatively: Table 340 2 provides the regression coefficients obtained by fitting data in Figures Sa and Sb to nonlinear 341 exponential regression equations of the form: 342 y = a*exp(-log(Ki,)/b) (9) 343 using the Leven berg-Marquardt algorithm for optimization. For application to a specific site, linear 344 interpolation of the values for a and bin Table 2 may be used to compute the height of capture in 345 the gravel pack and the proportion of flow originating from the gravel pack above the well screen for 346 values of the anisotropy ratio and of Kg other than those given in the Table. This modified analytical 347 tool provides a much more realistic source area than the much simpler graphical method employed 348 in many states as part of their source water assessment programs (e.g., California OHS 1999). 19 Hom a) b) E 30 0 !;I ~ 60 0 Kg:50 .,,_ 0 Kg: 125 0 25 50 m <> Kg: 250 gi_ .; 0 • A Kg: 500 ~ 20 C • Kg: 750 ◊ 40 A c Kg: 1000 en ~ a, • ~ C £ 15 ~ 30 0 • . 5 A m E 0 D '.§ A • -i 10 20 0 ◊ A " .: 0 • E 0 g 0 ◊ A ::, D 0 E 5 A 10 0 ◊ 0 "ij ~ Q 0 0 0 E Iii 0 0 0 1 3 5 10 30 100 300 3 5 10 30 100 300 349 Kh [m/d] Kh [m/d] 350 Figure 8: (a) Maximum virtual water particle heights in the gravel pack above the well screen 351 serving to capture water (b) Percentage of inflow into the well screen flowing through the gravel 352 pack from above the screen. Both for an anisotropy ratio of 2. 353 354 355 356 Anisotrop Kg Parameter for Adjusted Parameter for inflow Adjusted y maximum heights: r2 from above: r2 a b A b 2 50 33.31 1.05 1.00 12.69 0.60 0.99 2 125 77.01 0.88 0.99 21.44 0.69 0.99 2 250 87.95 1.02 1.00 30.59 0.76 0.99 2 500 163.35 0.93 1.00 40.57 0.86 0.99 2 750 172.04 1.00 1.00 47.16 0.92 0.99 2 1000 226.70 0.96 0.94 52.10 0.96 0.98 5 50 79.01 0.78 0.95 14.94 0.61 1.00 5 125 98.28 0.93 0.97 25.14 0.70 0.99 5 250 216.33 0.80 0.93 35.57 0.77 0.99 5 500 276.67 0.87 0.99 47.59 0.85 0.99 5 750 396.12 0.85 0.99 55.54 0.91 0.99 5 1000 401.91 0.88 1.00 61.39 0.96 0.98 _,. . . 357 Table 2. Coefficients and adjusted coefficients of detemunahon (r) for the equanons descnbmg the 358 maximum heights of the capture zone in the gravel pack, and the inflow of water entering the well 359 from the gravel pack above the screen. 360 361 362 20 Hom 363 Discussion 364 For application to specific sites, Figure 7 provides a tool to estimate the additional source area due to 365 the gravel pack, when compared to the simple approximation (eq. 2). These results can also be 366 applied for conditions with smaller or larger recharge rates, R ', than the rate R = 0.669 m/a used in 367 our computations. For R' not equal to R, results shown in Figures 7-10 and expressed in the above 368 equation are looked up for a scaled hydraulic conductivity K' rather than for the actual hydraulic 369 conductivity K, where K' = K · R 1/R. This scaling procedure is approximate because it does not 370 simultaneously scale other parameters controlling the observed results, e.g., screen length and 371 pumping rate. However, for applications in unconsolidated sedimentary aquifers, this scaling 372 approach works well as the drawdown created by domestic wells is relatively small. For depths to 373 the top of the screen different from that used here, the simple geometric conceptual model outlined 374 in Figure 1 and expressed in eq. 2 provides a framework for adjusting the distance of the source area 375 from the well head. Equation 9 (with Table 2) can be used to estimate the fraction of flow 376 originating from the elongated part of the source area. 377 378 The numerical modeling shows the significant influence of the gravel pack on the source area of a 379 domestic well, particularly for lower permeable aquifers (horizontal hydraulic conductivities of less 380 than 10 mid). In highly permeable aquifers (relative to the recharge rate of0.669 m/year used in this 381 study), the analytical model (eqs. 2, 3) provides a relatively good approximation of the upgradient 382 and downgradient edge of the source area. Lower hydraulic conductivities lead to significantly 383 longer capture zones than predicted by the analytical model (eqs. 2-3). In our configuration of screen 384 length and gravel pack length, which represents an average domestic well construction for Central 385 California, the elongation due to the presence of a gravel pack constitutes up to 70 % of the total 21 Hom 386 length of the capture zone. The elongation is relatively narrow but higher gravel pack conductivities 387 lead to significant increases in that width. The width of the main capture zone, in turn, slightly 388 decreases at higher gravel pack conductivities. The greater the difference between hydraulic 389 conductivity of the aquifer and that of the gravel pack, the greater is the elongated part relative to the 390 total length of the capture zone. 391 392 For many contaminants, chemical or microbial, aquifer attenuation is a dynamic, time-dependent 393 process. Travel times for potential contaminants decrease approximately linearly with increased 394 gravel pack length above the well screen. This is due to the strong influence of recharge on vertical 395 downward displacement of water (and contaminants) and the relatively small influence that the 396 domestic well pumping exerts on the overall groundwater flow field. A linear decrease in travel time 397 from the time of recharge until arrival at the gravel pack is associated with exponentially increased 398 contaminant concentrations. The gravel pack itself typically provides much less attenuation capacity 399 than the aquifer material. Hence, a short seal and vertically extended gravel pack constitute a 400 potential short-circuit for contaminants. 401 402 We also note that the fraction of flow captured by the gravel pack above the well screen may be 403 relatively small in a productive (high K) aquifer. But for some contaminants the resulting dilution 404 with (good) groundwater collected by the well at the depth of the screen may not be sufficient. This 405 includes contaminants that reach the water table at concentrations that are several orders of 406 magnitude above regulatory drinking water limits including solvents, pesticides, other organic 407 chemicals, and pathogens. A possibly common source of such contamination are septic tank leach 22 Hom 408 fields, which -in rural and semi-rural housing developments -are often located in the vicinity of 409 domestic wells. 410 411 Conclusions 412 Our work provides a tool to quickly estimate the size and location of the source area of domestic 413 wells in regions with significant recharge (for example, due to irrigation). The influence of the 414 gravel (or sand) pack in the well annulus above the well screen is explicitly accounted for. Results 415 allow for estimation of source area and gravel pack impact for a wide range of scenarios. 416 Importantly, we show that the gravel pack above the well screen poses a significantly increased risk 417 for domestic well contamination. A gravel pack that extends significantly above the well screen ( due 418 to short seal length), may significantly enhance the length of the source area, thus exposing the well 419 to a larger cross-section of potential contaminant sources. The extended gravel pack also decreases 420 travel time and distance for contaminants from the source area to the well allowing for contaminants 421 to partially circumvent natural aquifer attenuation. This is especially true in aquifers with low to 422 intermediate hydraulic conductivity (K :<::: 10 mid). We therefore strongly recommend that the gravel 423 (or sand) pack not be extended more than a few meters above the well screen of a domestic well. 424 425 Acknowledgments. We gratefully acknowledge the careful review and constructive comments of 426 Karen Burow, USGS, and two anonymous reviewers. Funding for this research was provided 427 through a fellowship of the German Academic Exchange Service (DAAD) to Judith Hom. 428 429 430 23 Hom 431 References 432 Belitz, K., and S. P. Phillips (I 995), Alternative to agriculture drains in California's San Joanquin 433 valley: results of a regional-scale hydrogeologic approach, Water Resour. Res., 31(8), 1845-1862. 434 435 Burow, K. R., J. L. Shelton, J. A. Hevesi, and G. S. Weissmann. 2004. Hydrogeologic 436 characterization of the Modesto area, San Joaquin Valley, California. USGS Scientific Investigations 437 Report 2004-5232. 438 439 California OHS. 1999. Drinking Water Source Assessment and Protection (DWSAP) Program. 440 Division of Drinking Water and Environmental Management, California Department of Health 441 Services. Sacramento, CA. http://www.dhs.ca.gov/ps/ddwem/dwsap/DWSAPindex.htm 442 443 Garg, S. P. and J. Lal, 1971. Rational design of well screens. J. Irrig. and Drain. Div., ASCE, 444 97(1):24-35. 445 446 Harter, T. 2008. Delineation of wellhead protection areas. In: T. Harter and L. Rollins (eds.), 2008. 447 Watersheds, Groundwater, and Drinking Water: A Practical Guide, University of California, UC 448 ANR Communications Services Publication 3497, Davis, CA 95616, 274p. 449 450 Houben, G. H. 2006. The influence of well hydraulics on the spatial distribution of well 451 incrustations. Ground Water 44, no. 4: 668-675. 452 453 McCray, J. E., S. L. Kirkland, R. L. Siegrist, and G. D. Thyne. 2005. Model parameters for 454 simulating fate and transport of on-site wastewater nutrients. Ground Water 43, no. 4: 628-629. 455 456 McDonald, M. G. and A. W. Harbaugh. 1988. Technics of water-resources investigation of the 457 United States Geological Survey. USGS Open-File Report 83-875. 458 459 Nahrgang, G., 1954. Zur Theorie des vollkommenen und unvollkommenen Brunnens. 43 p. 460 461 Phillips, S. P., C. T. Green, K. R. Burow, J. L. Shelton, and D. L. Rewis, 2007. Simulation of 462 ground-water flow in part of the Northeastern San Joaquin Valley, California, U.S. Geological 463 Survey, Scientific Investigations Report, SIR 2007-5009. 464 465 Pollock, D. W. 1994. User's guide for MODPATH/MODPATH-PLOT, Version 3: A particle 466 tracking post-processing package for MODFLOW, the U. S. Geological Survey finite-difference 467 ground-water flow model. USGS Open-File Report 94-464. 468 469 Todd, D. K. 1980. Ground water hydrology, 2nd ed. New York: John Wiley and Sons. 470 471 U.S. EPA. 1997. Response to Congress on use of decentralized wastewater treatment systems. 472 Washington, D.C.: Office of Water, U.S. EPA. 473 474 VonHofe, F. and 0. J. Helweg, 1998. Modeling well hydrodynamics. ASCE J. Hydr. Eng. 475 124(12):1198-1202. 24 Hom APPENDIXB Statistical Analysis Relating Well Yield to Construction Practices and Siting of Wells in the Piedmont and Blue Ridge Provinces of North Carolina United States Geological Survey Water-Supply Paper 2341-A Prepared in cooperation with the North Carolina Department of Natural Resource~ and Community Development SELECTED SERIES OF U.S. GEOLOGICAL SURVEY PUBLICATIONS Period lea ls Earthquakes & Volcanoes (issued bimonthly). Preliminary DetermlnaUon or Eplcenters(usued monthly). Technical Books and Reports Proresslonal Papen arc mainly comprehensive scientific reports of wide and lastins interest and importance to professional scientislS and en- gineers. Included are reports on the resu!IS of resource •tudies and of topographic, hydrologic, and geologic investigations. They also include collections of relalod papen addrusing diffem,t aspe<:ISof a single scien- tific topic. BuDetlns contain •ignificant dBla and inte,prelations th8l are of last- ing scientific interest but arc generally more limited in scope or geographic coverage than Professional Pupers. They include the results of resource studies and of geologic and topographic investigations: as well as collections of short papers rclau:d to a specific topic. Water-Supply Papers arc comprehensive reports thBI present sig- nificmt interpretive results of hydrologic investigations of wide interest to professional geologists, hydrologists, and engineers. The serics coven investigations in all phases of hydrology, including hydrogeology, availability of water, quality of water, and use of WIiler. Cliculan present adminiJtralive information or important scientific infonnarion of wide popular intttest in a format designed for dislribution at no cost to the public. Information is usually of shon--tcnn interest. Water-Resources Investigations Reports are papen of an inte,pre- tive nature made available to the public outside the formal USGS publi- cations .. rie,. Copies are n:pro<iu=I on reque.t unlil:e fonnal USGS publications, and they ue also av ail able for public inspection at depositories indicalod in USGS catalog,. Open-FIie Reports include unpublished mmruscript reports, maps, and other material tha1 arc made available for public consuliation at depositories. They arc a nonpermancnl form of publication that may be. cited in other publications as sources of information. Maps Geologlc Quadrangle Maps are multicolor geologic rnaps on topographic bases in 7 Ill.-or I 5-minutequadrangle fonnBIS (scales main- ly I :24,000 or I :62,500) showing bedrock, surficial, or engineering geol- ogy. Mops generally include brief texlS; some mops include •tructure and columnar sections only. Geophysical Investigations Maps are on topographic or planimelric bases at vmious scales; they show results of rurveys using geophysical techniques. such as gravity, magnetic, seismic, or radioactivity, which reflect subsurface structures that are of coonomic or ge.ologic significance. Ma."ly maps include correlatiom with the geology. Miscellaneous lnVl!Stlgallons Serles Maps arc on planimelric or topographic bases of regular and irregular arcss at various scales; they present a wide variety of fonnat and subject matter. The series also in- cludes 71/2-minutequadrangle photogeologic maps on planimetric bases which show geology u interpreted &om aerial photogrophs. Series also includes maps of Man: and lhe Moon. Coal Investlj:aUons Maps are geologic maps on topographic or planimetric: bases at various scales showing bedrock or surficia1 geol- ogy, stratigraphy. and structural relations in certain coal-resource areas. OU and Gas Investigations Charts show sll'llligraphic information for certain oil and gas fields and other areas having petroleum potential. Mlscellnneous Field Studies Maps are multicolor or black-and- white mops on topographic or planimenic bases on quadrangle or ir- regular areas a1 various scales. Pre-1971 maps show bedrock geology in relation to specific mining or mineral-deposit problems; post-1971 maps are primarily blade-and-white maps on various ,ubjectJl such as environmental studies or wilderness mineral investigations. Hydrologlc Investigations Atlases are multicolored or black-and- whitemops on topographicorplanimelric bases presenting a wide range of geohydrologic dalaofboth regular and irregular areas; principal scale is 1:24,000 and regional studies are at 1:250,000 scale or smaller. catalogs Permanent catalogs. u well as some othen, giving cornpre.hc:n- sive listing, of U.S. Geological Survey publications ore available under theconditions indicalod below &om the U.S. Geological Survey, Boob and Open-File Reports Section, Federal Center, Box 25425, Denver, CO 80225. (See larat Price and Availability List) "Publications of the Geologk:al Survey, 11179-1961" may be pur- chased by mail and over the counu:r in paperback book form and as a set of rniaofiche. "Publications of the Geologlcal Survey, 1962-1970" may be pur- chased by mail and ovu the counter in pape,back book form and as a Bel of microfiche. • PubUcaUons or the U.S. Geologic al Survey, 1971-1981" may be purchased by mail and over the counu:r in paperback book form (two volumes, publications listing and index) and as a set of microfiche. Supplements for 1982, 1983, 1984, 1985, 1986, and for subsequent years since the last pemwtc:nt catalog may be purchased by mail and over the counter in paperback book form. State catalogs, "List of U.S. Geological Survey Geologic and W111er-Supply Reporu and Maps For (Stale)," may be pun:hased by mail and over the counter in paperback booklet form only. "Price and A nllabWty List of U.S. Geologlcal Survey Publlca- tlons, • issued annually, is available free of charge in paperbeck book- let form only. Seltttod roplesor a monthly catalog "New Publications of the U.S. Geological Survey" available free of charge by mail or may be obtained over the counter in paperback book.let fonn only. Those wishing a free ,uhsaiption to the monthly catalog "New Publications of the U.S. Geological Survey" should write to the U.S. Geological Survey, 582 National Center, Reston, VA 220!/2. Note.-Price.5 of Government publications listed in older catalogs, announcements, and publications may be incorrect Therefore, the pri~ charged may differ from the prices in calalogs, announcemenlS, and poblicatlons. Chapter A Statistical Analysis Relating Well Yield to Construction Practices and Siting of Wells in the Piedmont and Blue Ridge Provinces of North Carolina By CHARLES C. DANIEL Ill Prepared in cooperation with the North Carolina Department of Natural Resources and Community Development A statistical analysis of data from more than 6,200 water ~ells was made to identify geologic, topographic, and construction factors associated with high-yield wells U.S. GEOLOGICAL SURVEY WATER-SUPPLY PAPER 2341 GROUND-WATER RESOURCES OF THE PIEDMONT-BLUE RIDGE PROVINCES OF NORTH CAROLINA DEPARTMENT OF THE INTERIOR MANUEL LUJAN, Jr., Secretary U.S. GEOLOGICAL SURVEY Dallas L. Peck, Director Any use of trade, product, or firm names in this publication is for descriptive purposes only and does not imply endorsement by the U.S. Government UNITED STATES GOVERNMENT PRINTING OFFICE : 1989 For sale by the Books and Open-File Reports Section, U.S. Geological Survey, Federal Center, Box 25425, Denver, CO 80225 Library of Congress Cataloging in Publication Data Daniel, Charles C., Ill Statistical analysis relating well yield to construction practices and siting of wells in the Piedmont and Blue Ridge provinces of North Carolina. (U.S. Geological Survey water•supply paper; 2341-A) Supt. of Docs. no. : I 19.13:2341A) "'Prepared in cooperation with the North Carolina Department of Natural Resources and Community Development."' Bibliography: p 1. Water, Underground-North Carolina-Statistical methods. 2. Wells-North Carolina-Statistical methods. 3. Water•supply-North Carolina-Statistical methods. I. North Carolina. Dept. of Natural Resources and Community Development. II. Title. Ill. Series. TD224.N8D36 1989 628.1"14'09756 88-600118 CONTENTS Abstract Al Introduction A 1 Purpose and Scope A3 Previous Investigations A3 Description of the Study Area A3 Physiography A3 Geology A6 Hydrogeologic Units A6 Geologic Belts and Terranes AB Compilation of the Data Base and Statistical Procedures AB lnfomtation Categories in the Data Base AB Statistical Procedures Al3 Relation of Well Yield to Construction Practices and Siting of Wells Al4 Results of the Analysis Al4 Well Yields by Hydrogeologic Unit A20 Well Yields by Geologic Belts and Terrones A22 Summary and Conclusions A23 References A26 FIGURES I. Map of North Carolina showing area of investigation. counties, and physio- graphic provinces Al 2. Index map of North Carolina showing study areas of reconnaissance ground- water investigations that were the sources of well data for this study A4 3. Map and geologic section showing the physical setting of the ground- water system in North Carolina AS 4. Index map of North Carolina showing areas of geologic maps used in compi- lation of the hydrogeologic unit map of the Piedmont and Blue Ridge provinces A9 5. Hydrogeologic unit map of Guilford and Alamance Counties and vicinity in the north-central Piedmont of North Carolina AlO 6. Map showing geologic belts, terranes, and some major structural features within the Piedmont and Blue Ridge provinces of North Carolina All 7. Graph showing variation of average yield. average depth. and average yield per foot of well depth with well bore diameter Al 7 8-IO. Contour plots of trend surfaces showing: 8. Relation between well yield, total well depth, and well diameter for wells that are located in draws and valleys Al8 9. Relation between well yield, total well depth. and well diameter for wells that are located on slopes and flats Al9 JO. Relation between well yield, total well depth, and well diameter for wells that are located on hills and ridges A20 11 . Graph showing variation of average yield and yield per foot of well depth with depth for wells having diameters between 5.5 and 6.5 in. All Content! Ill 12. Contour plot of trend surface showing relation between yield per foot of well depth, total well depth, and well diameter A22 13--14. Graphs showing: I 3. Average yield of wells of average construction in the hydrogeologic units of the Piedmont and Blue Ridge provinces of North Carolina A24 14. Average yield of wells of average construction in the geologic belts and tcrrancs of the Piedmont and Blue Ridge provinces of North Carolina A2S TABLES 1. Classification and lithologic description of hydrogcologic units in the Piedmont and Blue Ridge provinces of North Carolina A 7 2. Geologic belts and terrancs of the Blue Ridge, Piedmont, and Coastal Plain provinces of North Carolina A12 3. Total number of entries for each variable in the water-well data base A13 4. Average and median values of selected well characteristics according to topographic sening compared to statistics for all wells A14 5. Summary statistics defining depth to water, casing depth, and saturated thickness of regolith according to topographic group in the Blue Ridge and Piedmont physiographic provinces A16 6. Relation of selected well characteristics to the use of the well A16 7. Relation of well yields to hydrogeologic unit and topography A23 METRIC CONVERSION FACTORS For readers who wish to convert measurements from the inch-pound system of units to the metric system of units. the conversion factors are listed below: Multiply inch-pound unit By To obtain SI unit l..tngth inch (in.) 2.5.4 millimeter (mm) foot (ft) .3048 meter (m) mile (mi) 1.609 kilometer (km) Area square mile (mi2) 2.590 SQWU'C kilometer (km2) Volume gallon (gal) 3.785 liter (L) .003785 cubic meter (m3) Flow gallon per minute (gaVmin) 3.785 liter per minute (Umin) .003785 cubic meter per minute (m3/min) Flow per Length gallon per minute per foot 12.418 liter per minute per meter [(gaVmin)lft] [(Umin)/m] .01242 cubic meter per minute per meter [(m3/min)lm] ALTITUDE DATUM Sea level: ln this report "sea level" refers to the National GeodeticVertical Datum of 1929 (NOYD of 1929)-a geodetic datum derived from a general adjustment of the first-order level nets of both the United States and Canada, formerly called Sea Level Datum of 1929. IV Contenb Statistical Analysis Relating Well Yield to Construction Practices and Siting of Wells in the Piedmont and Blue Ridge Provinces of North Carolina By Charles C. Daniel Ill Abstract A statistical analysis was made of data from more than 6,200 water wells drilled into the fractured.crystalline rocks of the Piedmoiil:"Blue Ridge, and western ed~ of 'tlie C~astal Plain where crystalline rocks underlie sedi- ments at shallow depths. The study area encompassed 6.5 counties in western North Carolina, an area of 30,544 square miles, which comprises nearly two-thirds of the State. Additional water supplies will be needed in western North Carolina as population and industrial development continue to increase. Ground water is an attractive alter- native to surface-water sources for moderate to large supplies. The statistical analysis was made to identify geologic, topograptii~d·construction·fa'ctors=thar'are asrociated'wiikliigh:yield wells. =tfls-generally believed-tliat the crystalline rocks of the Piedmont and Blue Ridge provinces yield only small amounts of water to wells, that water is obtained from vertical fractures that pinch out at a depth of about 300 feet because of lithostatic pressure, and that the function of a large diameter well is primarily for storage. These concepts are reasonable when based upon the fact that the average well drilled in these rocks is a domestic well, 125 feet deep, 6 inches or less in diameter, and located on a hill or ridge. However, statistical analysis shows that wells in draws or valleys have average yields three times those of wells on hills and ridges. Wells in the most productive hydrogeologic units have average yields twice those of wells in the least productive units. Wells in draws and valleys in the most productive units average five times more yield than wells on hills and ridges in the least productive units. Well diameter can have a significant influence on yield: for a given depth, yield is directly proportional to well diameter. Maximum well yields are obtained from much greater depths than previously believed. For exam- ple, the average yield of &-inch diameter wells located in draws and valleys can be expected to reach a maximum of about 45 gallons per minute at depths of 500 to 525 feet; for similarly located 12•inch diameter wells, the average yield can be expected to reach a maximum of about 150 gallons per minute at depths of 700 to 800 feet. INTRODUCTION Additional water supplies will be needed in the Piedmont and Blue Ridge provinces of North Carolina (fig. I) as population and industrial development continue to increase. Municipal and industrial water supplies are derived almost exclusively from surface water sources. However. the potential for further development of surface water is limited, and ground water is an attractive alterna- tive for moderate to large water supplies. Ground water has many attractive features as a source of supply. Ground water in the crystalline rocks of the Piedmont and Blue Ridge provinces has a relatively low cost of development (Cederntrom, 1972).Generally, ground water in these areas is of good chemical quality and requires linle treatment. Because of the large quantity of water in storage, the ground-water system usually can sustain mod- erate yields during seasonal dry periods. The use of ground water generally pennits other land-use activities if they do not impede the infiltration of recharge or diminish water quality. The crystalline rocks that underlie the Piedmont and Blue Ridge are reputed to furnish only small quantities of ground water. This impression is the outgrowth of drilling large numbers of domestic wells that do not represent efforts to obtain quantities of water beyond the minimum requirement of 2 to 10 gallons per minute (gal/min). About 70 percent of all wells drilled in the Piedmont and Blue Ridge are for domestic supply. and most were located and drilled without regard to geology .topography, and optimal constrUction. In spite of these shortcomings, a significant number of wells yield a few tens to a few hundreds of gallons per minute. Additional high-yield wells likely can be developed at carefully selected sites throughout the area. Analysis Relating Well Yield to ConstrucOon and Sit;ng, Piedmont•Blue Ridge Provinces, North Carolina A 1 ••• . ,. 36° ... North Carolina study area: 30,544 square miles 65 counties , .. 1980 population 4.36 million Figure 1. Area of investigation showing counties and physiographic provinces. rs• 11• 0 25 50 100 MIL.ES •--~~--~~----~ 0 •• 100 KILOMETERS Results of studies in several areas of the Piedmont, both within and outside North Carolina, show that the ground-water system can suppon large well yields. For example, Daniel and Sharpless (1983) reported finding more than 300 wells in an eight-county area of central North Carolina that produce 50 gaVmin or more. Cressler and others (1983) found a substantial number of wells in the Georgia Piedmont that yield more than 100 gal/min and- some that yield nearly 500 gal/min. They also found 66 mainly industrial and municipal wells that had been in use for periods of 12 to more than 30 years without experienc- ing declining yields. Similarly, Cederstrom (1972) found that yields of I 00 to 300 gal/min are not uncommon for bedrock wells in the Piedmont and Blue Ridge provinces from Maine to Virginia. To evaluate the potential for large ground-water supplies in the Piedmont and Blue Ridge provinces of North Carolina, the U.S. Geological Survey-in cooperation with the North Carolina Department of Natural Resources and Community Development-conducted a 5-year study of ground-water resources in the region. This report is part of that study. Purpose and Scope This report describes a statistical analysis of data from a large number of water wells in the Piedmont and Blue Ridge provinces of North Carolina. The analysis was undertaken to identify factors that are associated with high-yield wells. The statistical analysis was made by using hydro- logic, geologic.topographic, and well-construction data that were obtained from records of more than 6,200 water wells. The wells are in an area including all of the Piedmont and Blue Ridge provinces in the State and an adjoining narrow strip at the western edge of the Coastal Plain province where a number of wells draw water from Piedmont crystalline rocks at shallow depth beneath the sedimentary cover. The study area encompassed 65 counties in North Carolina, an area of 30,544 square miles (mi2), which comprises nearly two-thirds of the State (fig. I). The records of water wells, obtained from published sources. were used to compile information on well yields and water levels; use of the water; well-construction vari- ables such as total depth. diameter, and casing depth: and the siting of wells in relation to topography and geology. A total of 14 geologic terranes considered to be hydrologically significant were identified in the study area. Within these terranes are 21 major rock types of igneous, metaigneous, metasedimentary, metavolcanic, and sedimentary origin that are considered to have quantifiable hydrogeologic properties. Because of their hydrogeologic properties, these major rock types are designated herein as hydrogeologic units. The data on both geologic terranes and hydrogeologic units were obtained largely from the work, both published and unpublished, of other investigators. Field studies were kept to a minimum. Previous Investigations Between 1946 and 1971, a total of 14 reconnaissance ground-water investigations (fig. 2) were completed that provided information on ground-water resources in all the counties in the Piedmont and Blue Ridge provinces of North Carolina. All but one of these reports (Peace and Link,1971) were prepared by the U.S. Geological Survey in cooperation with various Nonh Carolina State agencies. Included in the 14 reports, which were the main sources of data for this report. are maps showing well locations in each county and tables of well records providing details of well construction, yield, use, topographic setting, water-bearing formation·. plus miscellaneous notes. DESCRIPTION OF THE STUDY AREA Physiography North Carolina lies in three physiographic provinces of the southeastern United States (fig. 3): the Blue Ridge. the Piedmont, and the Coastal Plain (Fenneman, 1938). The Blue Ridge province in western North Carolina contains the greatest mountain masses, highest altitudes. and the most rugged topography in eastern North America. The province is marked by steep, forest-covered slopes that are cut by numerous small stream valleys. More than 40 peaks are greater than 6.000 feet (fl) in altitude and another 82 peaks are between 5,000 and 6,000 fl in altitude (Conrad and others, 1975). The province is bounded on the west in Tennessee by the Ridge and Valley province. On the east. the boundary of the Blue Ridge with the Piedmont province is marked by the escarpment of the Blue Ridge front-a prominent topographic feature thought in part to be associ- ated with faulting. The Blue Ridge front rises more than 1,700 ft above the Piedmont surface Ill the North Carolina- Virginia border and reaches a maximum relief of nearly 2,500 ft in central North Carolina. The topography of the Piedmont consists of low, well-rounded hills and long, rolling. northeast-trending ridges. The tops of many ridges and interstream divides are relatively !lat. They are thought lo be remnants of the Piedmont peneplain, an ancient erosional surface of low relief. More recent erosion and downculling by streams has dissected the Piedmont peneplain and created a local topo- graphic relief of 100 to 200 ft between interstream divides and stream bottoms. The Piedmont surface is 300 to 600 ft Analysis Relating Well Yield to Construdion and Siting, Piedmont•Blue Ridge Provinces, North Carolina A3 83' 35° 3 Trapp, 1970 10 4 Sumsion and Laney, 1967 11 5 Peace and Link, 1971 12 6 LeGrand and Mundorfl, 1952 13 7 LeGrand, 1954 14 81' so• Bain, 1966 May and Thomas, 1988 Schipl, 1961 Pusey, 1980 Mundorff, 1946 _ _L 78° 0 ' - 0 Figure 2. Study areas of reconnaissance ground-water investigations that were the sources of well data for this study. 76° " so 100 MILES so 100 KILOMETER$ ••• , .. ,.. FEET 8,000 SEA LEVEL 8,000 Figure 3. ••• ••• A BLUE RIDGE ... -,·· • •• -r· PIEDMONT Blue Ridge Front / Gneiss and schist Piedmont Charlotte Belt Plateau 790 1&• 11_• ______ c',•c•-----, ----~----·.--, .. _,. Triassic basin I Fall Line •~--""---'~'------•"OO MILES 0 ,. 100 KILOMETER'S COASTAL PLAIN Cape Hatteras Pamlico Sound A' Physical setting of the ground-water system in North Carolina (modified from Heath, 1980). in altitude along the eastern border and rises gradually to the west to about 1,500 ft in altitude at the foot of the Blue Ridge front. Scattered across the rolling Piedmont surface are remnants of once higher mountains that because of their resistance to erosion stand as much as 500 to 1,600 ft above the local land surface. Some form prominent lines of hills. Others are isolated hills and mountains. called monadnocks, that stand alone above the Piedmont surface and. although more common in the western Piedmont. are found through- out the province. The Piedmont is bounded on the east by the Fall Line where the hard crystalline rocks of the Piedmont give way to the softer sedimentary rocks of the Coastal Plain prov- ince. At the Fall Line. the swift-flowing streams of the Piedmont enter the Coastal Plain over a zone of rapids and low falls. The Coastal Plain has little relief in contrast to the adjoining Piedmont. It is marked by sluggish streams flowing in broad valleys cut into predominantly sand and clay units that thicken seaward from a feather edge at the Fall Line. Along the western edge of the Coastal Plain, the sediments are underlain at shallow depth by crystalline Piedmont rocks (fig. 3). Geology The geology of the Piedmont and Blue Ridge is extremely complex. All major classes of rocks-metamor- phic, igneous, and sedimentary-are represented. although metamorphic rocks are the most abundant. The metamor- phic and igneous rocks range in composition from felsic to ultrarnalic and range in age from Precambrian in the Blue Ridge to Triassic and Jurassic in the Piedmont. The meta- morphism of the rocks varies in grade from low rank to high rank: that is, varying in degree of recrystallization and destruction of the original texture: many have been folded and refolded during multiple metamorphic and orogenic events. The rocks are broken and displaced by numerous faults and zones of shearing, some of which are many miles in length. Nearly everywhere are rock fractures without displacement called joints. The joints commonly cluster in groups orientated about one or more preferred directions. Within the crystalline rocks of the Piedmont are down- faulted basins (grabens) filled with sedimentary rocks of Triassic age. Three or more periods of igneous intrusion (Fullagar, 1971) have resulted in the emplacement of plutonic bodies that range in size from batholiths down to dikes. sills. and veins. Most intrusions have been metamorphosed. deformed, and fractured, but some are massive and have little or no foliation. All rocks have been subjected to uplift, weathering. and erosion, which resulted in the widening of fractures and the formation of new openings such as stress-relief fractures. These breaks in the otherwise solid rock are the conduits for ground-water flow. All of the events and processes that are part of the geologic history of the area have given the hydrogeologic system properties that control the present-day movement and circulation of ground water. Bedding and planes of metamorphic foliation gener- ally are folded and tilted and can have almost any attitude and orientation. Fractures, bedding. and foliation create in- homogeneities in the rocks and result in permeability that is usually greatest parallel to bedding. foliation, and zones of fracture concentration: permeability is usually least at right angles to the plane of these features. Bedrock may be exposed at land surface on steep slopes, rugged hilltops. or in stream valleys. but nearly everywhere else it is overlain by unconsolidated material that may reach depths greater than 100 ft. Collectively this unconsolidated material, which is composed of saprolite, alluvium, and soil, is referred to as regolith. Saprolite is clay-rich, residual material derived from in-place weather- ing of the bedrock. When the bedrock weathers to form saprolite, the relict structures generally are retained, and the directional properties of permeability are also retained. In many valleys. the saprolite has been removed by erosion, and bedrock is exposed or thinly covered by alluvial deposits. Soil is present nearly everywhere as a thin mantle on top of both the saprolite and alluvium. The water-storing and transmitting characteristics of bedrock and regolith and the hydrologic relation between them determines the water- supply potential of the ground-water system in the Piedmont and Blue Ridge provinces. Hydrogeologic Units Within the Piedmont and Blue Ridge of North Caro- lina there are hundreds of rock units that have been defined and named by various.conventions more in keeping with classical geologic nomenclature than hydrologic terminolo- gy. The geologic nomenclature does little to reflect the water-bearing potential of the different units. To overcome this shortcoming and to reduce the number of rock units to the minimum necessary to reflect the differences in water- bearing potential, a classification scheme based on origin, composition, and texture was devised (table I). The ration- ale behind the hydrogeologic units shown in table I is the hypothesis that these factors would be linked not only to a rock's primary porosity but also to its susceptibility to the development of secondary porosity in the form of fractures and solution openings. The composition and texture would also determine, in part, the rate and depth of weathering of these units and the water-bearing properties of the resulting regolith. The origin of the hydrogeologic units is indicated by the rock class (igneous, metamorphic, or sedimentary) or A6 Ground-Water Resources of the Piedmont-Blue Ridge Provinces of North Carolina Table 1. Classification and lithologic description of hydrogeologic units in the Piedmont and Blue Ridge provinces of North Carolina Symbol Hydrogeologic unit IPI . .. . . . Igneous, felsic intrusive .... , ..... , ..... . III.. . . . . . Igneous, intermediate intrusive ......... . IMl ..... Igneous, mafic intrusive ............... . MIF.. .. . Metaigncous, felsic ................... .. Lithologlc description IGNEOUS INTRUSIVE ROCKS Light-colored. m~dy granitic rocks, fine-to come-grained, some prophyritic. usually massive. locally foliated; includes granite, granodioritc, quartz diorite. quartz monzonite. a1askites. Gray to greenish-gray, medium-to coarse-grained. massive rocks of dioritic composition; includes assemblages of closely associated diorite and gabbro where they are too closely associated to be mapped separately. Dark-greenish-gray to black, medium-to coarse-grained intrusive bodies: prima- rily gabbroic in composition. includes closely associated gabbro and diorite where they an, too closely associated to be mapped separately. ulrramafic rocks. diabase. dunite. MET AMORPIIIC ROCKS Metalpeous Rocks (lotrusln) Light-colored, massive to foliated metamorphosed bodies of varying assemblages of felsic intrusive rock types; local shearing and jointing arc common. MIi ..... Metaigneous, intermediate .•............ Gray to greenish.gray. medium-to coarse-grained, massive to foliated, well- jointed, metamorphosed bodies of dioritic composition. MIM .... Meta.igneous, mafic ..................... Massive to schistose greenstone, amphibolite, metagabbro and mctadiabase, may be strongly sheared and recrystallized; metamorphosed ultramafic bodies are often strongly foliated, altered to serpentine. talc. chlorite-trcmolite schist and gneiss. Metavol.canic Rocb (Extrusl~Eruptlvt) MVF .... Metavolcanic, felsic .................... Chiefly dense. fine-grained, light-colored to greenish-gray felsic tuft's and fclsic crystal tuffs. includes interbedded fclsic flows. Felsic lithic tuffs, tuff breccias, and some epiclastic rocks: recrystallized fine-grained groundmass contains feld- spar, sericite. chlorite. and quartz. Often with well-developed cleavage, may be locally sheared: phyllitic zones are common throughout the Carolina slate belt. MVI . . . . Metavolcanic, intennediate...... . . . . . . . . Gray to dark-grayish-green ruffs and crystaJ tuffs generally of andesitic composi- tion; most with well-developed cleavagei also includes interbedded lithic tuffs and flows of probable andcsitic and bmaltic composition and minor felsic vol- canic rocks. MVM . . . Metavolcanic, mafic...... . . . . . . . . . . . . . . Grayish-green to dark-green. fine-to medium-grained andcsitic to basaltic tuffs. crystal tuffs. crystal-lithic tuffs. tuff breccias and flows; pyroclastic varieties may contain lithic fragments; usually exhibits prominent cleavage; alteration minerals include chlorite, epidote, calcite. and trcmolite-actinolite. MVE.. . . Metavolcanic. cpiclastic......... . • . . . . . . Primarily coarse sediments including lnterbedded graywackes and arkoses and minor conglomerates, interbedded argillites and felsic volcanic rocks; much of the sequence is probably subaqueous in origin and most of the rocks were derived from volcanic tcrranes. MVU .... Mctavolcanic, undifferentiated ........... Volcanic rocks of all origins (extrusive and eruptive) and compositions (felsic to mafic) interbedded in such a complex assemblage that mapping of individual units is not practical. Metasedlmentary Rocks ARG . . . . Argillite .................... : . . . . . . . . . . Fine-grained, thinly laminated rock having prominent bedding plane and axial plane cleavage; locally includes beds of mud.stone, shale, thinly Jaminated silt- stone, conglomerate, and felsic volcanic rock. GNF .... Gneiss. felsic ............. , . . . . . . . . . . . . Mainly granitic gneiss; light-colored to gray, fine-to coarse-grained rocks. usu- ally with distinct layering and foliation, often interlayered with mafic gneisses and schists. GNM .... Gneiss, mafic .................•........ Mainly biotite hornblende gneiss; fine-to coarse-grained, dark-gray to green to black rock, commonJy with distinct layering and foliation. often interlayered with biotite and hornblende gneisses and schists. and occasional amphibolite layers. MBL. . . . Marble .......... , . . . . . . . . . . . . . . . . . . . . . Fine-to medium-grained, recrystallized limestone and dolostone: found prima- rily in the Blue Ridge belt. PHL.. . . . Phyllite.. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Light-gray to greenish-gray to white, fine-grained rock having well-developed cleavage; composed primarily of sericite but may contain chlorite: phyllitic zones arc common throughout the Carolina slate belt and probably represent zones of shearing alt.uough displacement of units is usually not recognizable. Analysis Relating Well Yield to Construction and Siting, Piedmont-Blue Ridge Provinces, North Carolina A7 Table 1. Classification and lithologic description of hydrogeologic units in the Piedmont and Blue Ridge provinces of North Carolina-Continued Symbol Hydrogeologic unit lithologic description QTZ .. .. Quartzile ............................. . Me1asands1one, often feldspalhic 10 highly feldspall>ic. ll>in-10 !hick-bedded with occasional graded bedding, includes meta•arkose and metaconglomerate; often interbedded with mica schist, phyllite, and slate. SCH . . . . Schis1 ................................ . Schistose rocks containing primarily the micas muscovite or biotite or both. occasional sericite and chlorite schists; locally interlayered with hornblende gneiss and schist. commonly with distinct layering and foliation. SLT ..... Stale ................................. . Fine.grained metamorphic rock formed from such rocks as shale and volcanic ash. possesses the property to part along planes independent of the original bed· din (slaiy cleavage). MISCELLANEOUS TRI . . . . . Triassic sedimentary rocks ............. . Mainly red beds. composed of shale, sandstone, arkose. and conglomerate (fan• glomerate near basin margins). CPL. . . . . Coastal Plain basement ................ . Undifferentiated crystalline ba.~ment rocks of igneous and metamorphic origin overlain unconfonnably by sedimentary sands. gravels, clays, and marine deposits. subclass (melaigneous, melavolcanic. or metasedimenlary). The composition of the igneous, metaigneous, and meta- volcanic rocks is designated as felsic, intennediate.or mafic excepl for lhe addition in the me1avolcanic group of epiclastic rocks and compositionally undifferenliated rocks. These las1 two groups were necessary because of lhe significant areas of epiclas1ic rocks where reworking by sedimenlary processes and admixlure of terrigenous sedi- ment during deposition made lhe rocks lexturally distinc1 and the other areas where the complex and small-scale stratigraphic changes made differentiation of separale units impractical. Composition is also shown in lhe me1asedi- mentary units of gneiss, marble, and quartzite. The other melasediments are designated primarily on lhe basis of texture (grain size, degree of metamorphism, and develop- menl of foliation). The two miscellaneous classifications account for the sedimeniary rocks within 1he Triassic basins and 1he undif- ferentialed crys1alline basement rocks east of 1he Fall Line that are overlain unconfonnably by sediments of Crelaceous age and younger. By using the classification scheme in table I and the most recent geologic maps available (fig. 4), a hydrogeo- logic unit map was compiled for the study area. Part of this map for Guilford and Alamance Counties in lhe north- central Piedmont (fig. I) is shown in figure 5. Well-localion maps were later superimposed on lhis hydrogeologic unit map, and the units corresponding to lhe well locations were coded and entered inlo a computerized dala file for analysis 10 detennine the well yields in each unit. Geologic Belts and Terranes The Piedmont and Blue Ridge have been divided inlo a number of northeast-ttending geologic bells (fig. 6). Within a belt, rocks are to some degree similar to each other wi1h respect to general appearance, melamorphic rank, structural history, and relative abundance of igneous, metaigneous, metasedimeniary. and metavokanic rocks (Butler and Ragland, 1969). Areally, the most significant are lhe Blue Ridge, Inner Piedmonl, Charlolle. Carolina sla1e, and Raleigh belts. Two geologic terranes importan1 10 Ibis study have been added to lhe generally recognized belts. These are lhe Triassic basins and lhe Coaslal Plain immedia1ely eaSI of lhe Fall Line. A brief summary of lhe belts and lhe hydrogeologic units 1ha1 conSlitute the belts is given in table 2. Wells lapping rocks wilhin 1hese bells and terranes were analyzed to determine well yields within each area. COMPILATION OF THE DATA BASE AND STATISTICAL PROCEDURES Information on 6,224 wells was compiled from pub- lished sources (fig. 2) and s1atis1ically analyzed lo identify relations between well yield and various geologic, lopo- graphic, and construction faclors. This compilation con- lained well records from every coumy in lhe 65-counly study area and included 4 I 9 wells that derive water from crys1alline rocks buried benealh the 1hin sedimen1ary cover along the western edge of lhe Coastal Plain (fig. 3 ). Information Categories in the Data Base Specific types of infonnation calegories (variables) in the data base included (I) the county where the well is located, (2) the published well number. (3) 1he 101al depth of the well, (4) well diameter, (5) casing depth, (6) static wa1er level below land surface, (7) yield, (8) inlended use when drilled, (9) lhe 1opographic selling of lhe well sile, (10) the hydrogeologic unil into which lhe well is drilled, ( 11) lhe geologic bell or terrane in which lhe hydro geologic unit is found, and ( 12) 1he reference to 1he published report AB Ground•Water Resources of the Piedmont-Blue Ridge Pro\linces oi North Carolina ... .. a. b. C. d. e. I. g. h. ... 83' ... .,. ... 10• 78' 77• 78° -------.1----,------=---------- e a INDEX TO GEOLOGIC MAPS USED IN COMPILATION Compilation by C.C. Daniel Ill and Robert A. Payne, 1984·85, Based on sources indicated by letter as loUows: Stuckey, 1958 L McDaniel, 1980 Hadley and Nelson, t 971 ~ Wilson and Spence, 1979 Rankin and others. 1972 k. wnson, 1979 Goldsmith and others, 1982 I. WllsOn, t 981 Espenshade and others, 1975 m. Stromquist and Sundelius, 1975 Carpenter, 1982 n. Stromquist and others, 1971 Burt, 1981 0. Selders, 1981 Wilson and others, 1981 '\\ / ~ 0 0 50 100 MILES 0 50 100 KILOMETERS f',gure 4. Areas of geologic maps used in compilation of the hydrogeologic unit map of the Piedmont and Blue Ridge provinces. ~ .. Cl i! ~ ~ I ~ .. ! C ij ~ 5!. ;. .. [ 3 g 7 .. i: .. .. ;;: ":i .. i. ~ n i 0 ~ z 0 :,. ,,. n .. a if .. TN ... w "' -.. "' G F 0 R 0., "' u.. .., , .. --i/i ,,, I 1'1 ~-WI i ------ R A N D ""' "" 0 • ,,."r~t 0 • 10 KILOMETERS - -- "" D . <5f ~,I I 1,111 ~~, ... ~~; "" -, ... --~, , i --~ . , ' c:!fa --, M A N C ,.. 1!.llPLANATION HYDROGEOLOGIC UNITS IFI Igneous, felalc tnlrU&IY8 MIF Melatgneous, fe!sic MIi Melaigneous, intermediate MIM Melalgneous, maflc MVF Metavolcanic, lelaic MVI Metevotcanic, intermediate MVM Metavolcanic. mafic MVE Metavolcank:, eplclastlc ARG ArgDBta GNF Gneiss, felaic GNM Gneiss, mallc PHL Phyllite SCH Schist TRI Trlassic sedimentary rocks MAP OF NORTH CAROLINA G E "" Figure 5. Hydrogeologic unit map of Guilford and Alamance Counties and vicinity in the north-central Piedmont of North Carolina. ,.. ,.. ... .,. ,,. ~-- " Cl C 011 0 I A Bt•vard lault EXPLANATION IIIU •WUrohY bait SAUAATOWN MTN. ANT. -Saun11cwn wo1,1rnaln1 arrtlc:1/no,Lum SR • 8f!ll1h Rl'fltf dochthOl'I T..OCB -Oa'rie C~ty Trianle basin Tr-ORB • Dan River Trlaulc: ba1in cs -CaroMa st.w b.it CP • Cou1a• Plain Boundary betw••n Btua Ridge and Piedmont phy~aohlc orovlncea ..,....--r Thru11 faun with teeth on uplhrown bloek ... ... ,,. , .. ,,. , .. VIR01NIA PIEDMONT COASTAL PLAIN --------- 0 50 100 11111._(g 50 I 00 KII.OMETERS Figure 6. Geologic belts, terranes, and some major structural features within the Piedmont and Blue Ridge provinces of North Carolina (from Brown and Parker, 1985). Table 2. Geologic belts and terranes of the Blue Ridge, Piedmont, and Coastal Plain provinces of North Carolina (The hydrogcologic units an: described in lable I] Beh or terrane Murphy belt .......................... . Blue Ridge belt ...................... .. Qiauga belt. .......................... . (includes Brevard fault zone). Inner Piedmont belt. ................... . Smith River .................. , .. , ..... . allochthon. Sauratown Mountains .................. . anticlinorium. Kings Mountain belt. .................. . Qiarlotte belt ......................... . Milton belt. ........................... . letter designation MU ...... . BR ....... . CA ....... . IP ........ . SR ....... . SA ...... .. KM ....... CH ........ MI ........ Boundaries SU1TOunded by metasedimentacy rocks of Blue Ridge belt. Sedimentacy rocks of Ridge and Valley on north- west and Brevard fault zone on southeast. Blue Ridge belt on northwest. Inner Piedmont on southeast. Chauga and Blue Ridge belts on nonhwcst. Kings Mountain and Olarlotte belts on southeast. Blue Ridge belt on northeast and Sauratown Moun- tains anticlinorium on southeast. Smith River allochthon on northwest, Inner Pied- mont belt on southwest. and Dan River Triassic ba-.in and Milton belt on southeast. Inner Piedmont belt on northwest and Charlotte belt on southeast. Kings Mountain and Inner Piedmont belts on north- west. Milton belt on north, Gold Hill shear zone and Carolina slate belt on southwest. Igneous and metaigneous rocks of Charlotte belt on south, Carolina slate belt on southeast, Dan River Triassic basin and Sauratown MountaiIL'i anticli- norium on northwest. Dominant hydrogeer logic units SCH. SLT. MBL. ONF. ONM, SCH, QTZ. PHL. ONF. ONM. ONM. MIF. ONF. ONM. ONF. QTZ. SCH, MIF. ONF. Mil, MIF, MIM, JFI, MVU. ONM. ONF. Gold Hill shear zone................ . . . . GH... . . . . . Metavolcanic and metaigneous rocks of Charlotte PHL. belt on nonhwest and metavolcanic rocks of Cartr line slate belt on southeast. Carolina slate belt...................... CS . . . . . . . . Gold Hill. Charlotte, and Milton belts on north• ARO. MVE, MVU in southwestern half of belt-MVF, ARO, MVU, MIF. Mil in northeastern half of belt. west, Coastal Plain on southeast. Raleigh belt............................ RA........ Bordered by Carolina slate belt rocks on east and Mlf, ONF, SCH. west. Coastal Plain sediments on the south. Triassic basins . . . . . . . . . . . . . . . . . . . . . . . . . TR........ Several bodies of sedimentary rock downfaulted TRT. into the metamorphic crystalline rocks of the Pied- mont. Coastal Plain........................... CP .. .. . . .. Western edge of Coastal Plain province. CPL. from which the well record was obtained. The total number of entries for each variable is shown in table 3. For inclusion in the data base, a well had to satisfy certain requirements. The well had to be drilled into bedrock, and the yield and location had to be known. All wells in the resulting compilation are cased to the top of bedrock and have no screened or slotted intervals in the regolith, and nearly all are finished as open holes drilled into bedrock. A small number of wells included in the data base have casing. slotted casing. or screen extending into · bedrock to prevent fragmental rock debris from entering the well bore. An extreme example is a well that is 600 ft deep and is cased to the bottom of the hole. No other well has more than 300 ft of casing, and only 157 wells, or 2.5 percent, are cased to within the bottom 5 ft of the well. A12 Ground-Water Resour<ft of the Piedmont-Blue Ridge l'l'OviltCOI of North Carolina Table 3. Total number of entries for each variable in the water-well data base Variable Total number of data entries County . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.224 Well number.. . .. . . .. . . . . . .. . . . . . . . . .. . .. .. .. . .. 6.224 Total depth .. .. .. . . . .. . . . . . . .. . . .. .. .. .. .. .. . . .. 6.204 Well diameter.. .. .. .. .. . . . . . .. .. . . .. . . . . . .. . . .. . 6,0!iO Casing depth..... . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.038 Static water level. ........... ,.................... 3,130 Yield. .. .. .. .. .. .. .. . .. .. .. . .. .. . .. .. .. .. . .. .. .. 6,224 Use............................................ 6,205 Topographic setting.............................. 5,234 Hydrogcologic unit... .. .. . . . . .. .. .. .. . .. .. .. .. .. 6,224 Geologic belt .. .. .. .. .. .. .. . . . .. .. .. .. .. .. . .. .. . 6.224 Reference....................................... 6,224 The wells range in diameter from 1.25 to 15 inches (in.), and most (69 percent) of the wells have diameters between 5.5 and 6.5 in. Only two drilled wells were as large as 15 in. Large-diameter bored or dug wells were not included in the compilation because these wells arc not typical of modem well construction. Nearly all new wells in the Piedmont and Blue Ridge arc drilled by air rotary methods. Further, large-diameter wells arc rarely dug below the top of bedrock and do not represent attempts to obtain quantities of water beyond that necessary for domestic supplies. Transparencies were made of well-location maps given in the published sources (fig. 2) and overlaid on maps of the hydrogeologic units and geologic belts to assign the wells to the units and belts in which they occur. The hydrogeologic units reported in these publications were not entered into the data file because of the conflicting variety of names and naming conventions that were used by the many authors. The reported hydrogcologic units were not ignored, however. If a well was located on or near a contact between units used in this report, the published description helped guide the choice in the assignment of the unit and in some places pointed out the need for revisions to the hydrogeologic unit map. The published reports also were used to identify wells drilled into diabase dikes. Diabase dikes arc common in the Piedmont (Reinemund, 1955; Weigand and Ragland, 1970; Ragland and others, 1983), but generally they arc too narrow to accurately correlate with well locations at the scale of the maps being used. Wells drilled into diabase dikes arc included in the igneous, mafic intrusive (IM!) hydrogcologic unit. By using a combination of the new maps and the published descrip- tions, each well in the data base subsequently was assigned to I of the 21 hydrogeologic units. All data related to well construction, yield, topo- graphic setting, and static water level were entered as reported. The intended use of each well was inferred from the listed owner and other information in the remarks column of the well-record tables. Wells were placed in one of three use categories: domestic, commercial-industrial. and public supply. Domestic wells serve single family residences or, at most, a small number of homes. The commercial-industrial category includes wells that serve businesses ranging in size from large mills and factories down to service stations and small shops. Public-supply wells serve municipalities. subdivisions, trailer parks, hos- pitals, churches, campgrounds, and other facilities having a relatively large number of users. Every item of information was not available for every well. The static water level had the fewest number of data entries; water levels were reponcd for only slightly more than one-half of the wells. The second smallest number of entries was for casing depth; less than two-thirds of the well records had this information. The other variables had much more complete records. The effect of these incomplete records will be seen in the statistical analyses that follow. especially for computations that arc based on more than one variable. For example, in a calculation of yield per foot of well depth by topographic setting, the variables yield, depth, and topographic setting had 6,224, 6,204, and 5,234 data entries, respectively. Yet the final computation was based on the 5,221 wells for which all three items of information were available. This was generally the pattern; the final computation was based on no more than and commonly fewer observations than the smallest number of variable entries. Statistical Procedures The data were statistically analyzed by using pro- grams developed by the SAS Institute (SAS lnstinne, Inc., 1982a) that arc available on the U.S. Geological Survey computer system in Reston, Va. The most commonly used SAS procedures were SORT, UNIV ARIA TE, RSQUARE, GLM, and ANOV A. The SORT procedure (SAS Institute, Inc., 1982a) is a SAS utility procedure that sorts observations in a data set by one or more variables. In this study, the SORT proce- dure was used to sort the well data by topographic position, use, hydrogeologic unit, and geologic belt so that statistics could be computed for the soncd groups of data. The UNIVARIATE procedure (SAS Institute, Inc., 1982a) produces simple descriptive statistics including the mean, median, range, standard deviation, and quantiles for numeric variables. A SAS procedure called RSQUARE (SAS Institute, Inc., I 982b) was used for regression analysis because it allows many possible regressions to be fitted to the data and systematically analyzed to identify those combinations of variables that best explain the variation in the data. Those variables that repeatedly appeared in the models offering the highest r•squarc were further tested by using SAS procedure GLM (General Linear Models) (SAS Institute, Inc., Analysis Relating Well Yield to Construction and Siting, Piedmont-Blue Ridge Provinces, North Carolina A 13 1982b), which uses the method of least squares to determine regression coefficients, intercepts, and statistical properties of the models being tested. Analysis-of-variance tests using the procedure ANOV A (SAS Institute,Inc., 1982b) were made of the data in the topographic classifications, hydrogeologic unit clas- sifications, and geologic belt classifications to determine if any of the apparent differences, or Jack of differences, in mean values are statistically valid. Because the sample cells have unequal numbers of observations, Tukey's studentized range test, honestly significant difference (HSD) procedure (Steel and Torrie, 1960, p. 109-110), was used to make the multiple comparisons and to test for significant differences at the 0.95 confidence level. Unequal cell size was not the only reason for using Tu key's procedure. It is also a conservative test compared to other procedures such as Duncan's multiple-range test (Steel and Torrie, 1960. p. 107-109), which is most effective with samples of equal cell size, and controls for the experiment-wise error rate rather than on a percomparison basis. As a result. there is Jess chance that Tukey' s procedure will declare some differences between means to be significant even when the means are a homogeneous set. Duncan's multiple range test and the Duncan-Waller k-ratio t-test were also attempted on data sets that were manipulated to generate equal cell sizes. Equal cell sizes were generated by taking the percentile values of frequency distributions of data within a sample cell: this produced cells containing I 00 observations. This transformation woticed well for sample cells having large numbers of observations in a distribution that was not excessively skewed (skewness Jess than 4.0) and with similar values of skewness. When these two conditions were not met, the cell mean from the frequency distribution was different from the cell mean of the raw data. Because of this problem, the analysis-of-variance tests using Duncan's method and the Duncan-Waller method produced inconsistent results, although a pattern usually emerged that was similar to the results from Tukey's procedure. Because of the properties ofTukey's procedure, the nature of the data that were being tested, and for overall consistency, Tukey' s HSD procedure was used for all analysis-of-variance tests described in this report. Further discussion of analysis of variance, including Tukey's HSD procedure, can be found in Steel and Torrie (1960) and SAS Institute, lnc.(1982b). RELATION OF WELL YIELD TO CONSTRUCTION PRACTICES AND SITING OF WELLS Results of the Analysis The first group of statistics, presented in table 4, characterize the wells in the study area with regard to their physical and hydrologic characteristics. In the left half of the table, the average and median values of these character- istics are shown for wells in each of six topographic settings. The topographic settings are arranged in order of Table 4. Average and median values of selected well characteristics according to topographic setting compared to statistics for all wells Topographic setting All wells Well characteristic, Draw Valley Slope Flat Hill R;dgc Average First Median Third Ninth Number Average yield1 quartile guanlle dectHe of wells (gallons per minute) ......... 33.J 25.7 17.1 16.8 to.8 9.7 17.2 5 10 20 36 5,234 Median yield (gaUons per minute) ......... 20 IS 10 10 6 6 5,234 A vcrage yield per foot (gallons per minute per foot) .. .220 .205 .128 .131 .093 .086 .131 .038 .080 .165 .300 5,221 Median yield per foot (gallons per minute per foot) .. .154 .14.l .082 .083 .056 .058 5.221 A vcragc depth (feet) ...................... 175.1 157.8 152.6 150.0 150.2 153.1 154.0 85 119 179.5 297.4 5,221 Median depth (C..t) ...................... 134 104 118 119 117 112 5,221 Average casing (feel) ...................... 52.4 49.0 53.6 55.0 51.2 57.2 52.9 28 45 70 97 3,375 Median casing (f .. t) ...................... 46 40 47 50 43.5 42 3,375 Ave~ water level (feet low land surface) ..... 24.3 18.6 32 .. l 28.6 38.6 43.6 32.2 18 28 40 60 2,825 Median watcr level (feet below land surface) ..... 20 ,s 28 25 34 40 2,825 Average saruratcd thickness of rcgolith (feet) ............ 31.7 35.4 23.6 27.5 20.5 18.4 24.8 0 15 40 65 2,161 Median wuratcd thickness of regolith (feet) ............ 25 29 14 19 9 10.5 2,161 'Unadjusted for differences in depth and diameter. A 14 Ground-Waler Resources of the Piedmont-Blue Ridge Provinces of North Carolina decreasing average (mean) yield. The statistics of well characteristics in the six topographic settings can be com- pared to statistics computed for all wells in the sample that are given in the right half of the table, which defines the frequency at which a given value of a well characteristic can be expected to occw. At the first quanile, 25 percent of the wells in the sample have values that fall below the given value; at the median or second quartile. half the wells have values below the given value; at the third quanile, 75 percent of the wells fall below the given value; and at the ninth dectile. 90 percent of the wells are below the given value. The yield per foot of well depth and saturated thickness of regolith are computed characteristics. The yield per foot is the yield divided by the total depth of the well. The saturated thickness of regolith is the difference between the depth of casing and the depth of the static water level. If the water level in a well was below the bottom of the casing, the saturated regolith thickness of that well was considered to be zero. In the computation of the saturated thickness of regolith, casing depth was used to estimate regolith thick- ness. The depth of surface casing in a drilled well is a good approximation of regolith thickness in the Piedmont and Blue Ridge (Daniel and Sharpless, 1983; Snipes and others, 1983). Surface casing is usually set no more than I or 2 ft into fresh bedrock, just below the interface between it and the overlying regolith. Wells drilled in North Carolina since passage of the North Carolina Well Construction Act of 1967 (Heath and Coffield. 1970), however, are required to have a minimum of 20 ft of casing, regardless of how shallow the bedrock may be. Casing data from these wells can lead to overestimated regolith thickness. Fortunately, from a statistical standpoint, many of the records used in this study were for wells drilled prior to I 967. Records of casing depths as shallow as I ft for wells on bare-rock exposures are included in the data compilation. These data better reflect the natural range of depths to bedrock and thus provide for a more accurate approximation of regolith thickness. The data in table 4 show a general pattern of decreasing yield, yield per foot, and saturated thickness of regolith at higher topographic settings (ridges and hilltops). The depth to the water table follows the opposite pattern. The amount of casing and the well depth do not show any apparent relation to topographic setting e>cept that wells in draws average from 17 to 25 ft deeper than wells in other topographic positions. Analysis-of-variance tests of the data in the six topographic settings of table 4 were made in two steps, first on the data in the six settings and then on grouped data where significant differences were not found. In the first analysis, casing depth was not statistically different in any of the six topographic settings. The average depths for wells on slopes, flats, hills, and ridges were also statistically the same. The yield and depth of wells located in draws was statistically different (greater) from the yield and depth of wells located in valleys and other topographic settings. The remainder of the data tended to cluster in three topographic groups made up of those wells in draws and valleys, on slopes and flats, and on hills and ridges. It is important to point out that analysis-of-variance tests on yield per foot data indicate that wells in draws and valleys are statistically one group, because of adjustment of the yield to account for the differences in well depth in these two topographic settings. This finding is also an indication of the relation between well yield and well depth that will be described in more detail. In the second pan of the analysis. the data were merged according _to the three principal topographic groups identified in the first pan of the test. Analysis of variance on the grouped data still found no difference in casing depth, nor did well depths on slopes and flats differ from well depths on hills and ridges. Because the statistical tests showed that the yield per foot for wells in draws was the same as for wells in valleys, the yield and depth data for wells in the~-e settings were combined. 1ne remainder of the data fell into one of the three topographic groups and were statistically distinct from the other groupings for a given variable. Yields of wells in draws and valleys average nearly three times the yields of wells on hills and ridges. The highest yielding wells also were the wells having the greatest saturated thickness of regolith and the highest water level. Statistics showing the depth to the water table. casing depth, and saturated thickness of regolith for various topographic settings in the three physiographic provinces in the study area are given in table 5. The influence of topography on the depth to the water table is apparent. The effect of the higher relief and more rugged topography in the Blue Ridge is reflected by the greater depths to the water table than in comparable topographic settings in the Pied- mont. An unexpected finding is the similarity of the satwated thickness of regolith in the Piedmont and Blue Ridge. This may be due in pan to compensating conditions created by differences in rainfall and relief in the two provinces. Generally. there is more rainfall and more ground-water recharge in the Blue Ridge than in the Piedmont. But there also is greater relief, and presumably steeper ground-water gradients, in the Blue Ridge that results in greater ground-water discharge. Although there is less rainfall in the Piedmont (Eder and others. 1983), the lower relief results in lesser rates of ground-water dis- charge. Thus, the amount of ground water in long-term storage in the two provinces is roughly equal. Although the data for casing depth in table 4 indicate little difference between wells in different topographic settings when the study area is considered as a whole, the data in table 5 show that there is an increase in casing depth at higher topographic settings in the Blue Ridge. For wells in the Piedmont, there is no apparent relation between Analysis Relating Well Yield to Construction and Siting, Piedmont-Blue Ridge Pro\'inces, North Carolina A 1 s Table 5. Summary statistics defining depth to water, casing depth, and saturated thickness of regolith according to topographic group in the Blue Ridge and Piedmont physiographic provinces [Statistics for ~'Clls pcoctrating bedrock beneath the-wcs1cm ed~c of the Coastal Plain sediments arc gi\.·cn for comparison) elu@ Ridge Piedmont Coastal f'taln, Well characteriMic OraW!i Slopes Hills Number DraW!i Slopes HIiis All Number Number All All and and and wi!lls of wells and and and wells of wells welts of wells 'lalli::n """ ridges valle;r:s flats ridges A vcmgc water level (feet below land surface) ...... 23.4 37.5 6'.?.9 37.l 507 22.l 29.3 36.8 31.3 2.326 18.8 145 Median water level (feel below land surface) ..... 18 35 50 30 507 20 25 32 27 2,326 15 145 Average cn.c;i.ng (feet) ..........•...........• SO.I 57.7 66.6 56.8 698 52.7 53.2 50.0 52.0 2,684 71.7 293 Median casing 53.5 698 45 293 (feet) ....................... 43 55 60 46 41 44 2.685 63 A verasc saturated thickness or rtgolith (fttll .. , ....... 32.2 27.6 20.8 :?.8.0 422 33.6 '.?4.6 20.4 24.0 1,749 47.7 112 Median saturated thickness of rcgolith (fccu .......... 28 20 lO 20 421 2R 15 9 13 1,749 44.5 112 1Topogruphy of bedrock swface cannot be determined. Influence of topography on well yield in Coast~) Plain is unknown. Table 6. Relation of selected well characteristics to the use of the well [gal/min, galloru. per minutt: (gal/minl/ft. gallons per minute per foot: ft, feet] Percentages of wells according to use in selected topographic settings Statistical summary of well characteristics according to use Use of well om, Valll!y Slope Flat Hut C>omestic ......... , ...... 47.5 54.5 71.5 72.0 82.0 Commercial-industrial ...... 31.0 27.9 13.8 12.5 Public ................... 21.5 17.6 14.7 12.5 casing depth and topographic setting. This difference may be due to the greater relief in the Blue Ridge. In relation to use (table 6), more than one-half the wells in draws were commercial-industrial or public supply, and nearly one-half the wells in valleys were in the same two use categories. At the other topographic extreme, more than 80 percent of tbe wells on hills and ridges were domestic supply wells. The yields of domestic wells aver- age about one-third the yields of the commercial-industrial and public-supply wells and are about 100 ft shallower. lnfonnation on well diameter (not shown) also indicated that domestic-supply wells had the. smallest average diam- eters and public-supply wells had the largest. Fewer than 2 percent of domestic wells were 8 in. in diameter or larger. whereas 20 percent of the commercial-industrial and 26 percent of the public-supply wells were 8 in. or larger. The implication of the data in this table is that public-supply and commercial-industrial wells are more likely to be sited and constructed in an effort to obtain as much water as possible. whereas many domestic wells are at sites on hills and ridges selected for setting and view. Also, many secondary roads tend to follow the low ridgelines and drainage divides connecting the better drained agricultural land. and many rural homesites are near these roads. 7.8 10.2 Rktge Averle Average Average Average Average water Number yield yleldtfoot dept casing level of wells (gallminl (gallmln)lh (fll (ft) (ft) 83.6 I 1.6 0.117 123.6 51.8 30.8 4,408 3.3 27.7 .161 216.5 60.9 31.2 872 13.1 33.9 .171 229.8 69.2 34.7 905 The summary statistics strongly suggest a relation between well yield and well depth and diameter, a definite relation between topographic group and several well char- acteristics, including yield, and an apparent cultural bias in the siting and construction of wells related to the intended use of the well. The relation of well .yield to rock type, which has been described by many past authors, also may be distorted by cultural bias in siting and construction. For example. in the upper Cape Fear River basin, as described by Daniel and Sharpless ( 1983), the most productive rock unit is the mafic.volcanics unit. They showed a concentration of high-yield wells in central and northwestern Alamance County coinciding with the area underlain by the mafic- volcanics. Historically. this area has been a major center of textile manufacturing and has a number of factories and mills. The smaller towns have public water systems fur- nished by wells. and many of the mills have, or have had, their own ground-water supply systems. Thus, the area underlain by the mafic-volcanics unit may have appeared to be the most productive simply because it contained more large-diameter. deep wells than any other area in the basin. The relation between well yield and well depth and diameter is indicated in figure 7. where average yield, A 16 Ground•Water Resources of the Piedmont•Blue Ridge Pro\'inces of North Carolina (/) w ::c () ~ ~ 12 1 1 10 9 569 0.235 ::::s~ 482 0.197 ~l\:. ~Average depth In feet ~·i~~i~~ .... ••,•.· 8 321 0.206 a: o 148 '-------Average yield per loot of w 7 I-· well depth. In gallons w :; 6 per minute per foot :s 141 0.134 0 w 5 0.113 a: Average yield 17 .0 gallons per minute 0 4 [D 0.0112 Average yield per toot 0.131 gallons .J per minute per loot .J 3 130 0.076 w Average depth 151.9 feet ;!:; 2 121 0.065 Analysis based on data from 6,074 wells 1 81 0.092 0 20 40 60 80 100 120 140 160 180 AVERAGE YIELD, IN GALLONS PER MINUTE Figure 7. Variation of average yield, average depth, and average yield per foot of well depth with wellbore diameter. average depth, and average yield per foot of well depth are shown for wells of different diameters. 'The diameters arc subdivided into 1-in. intervals; the actual diameters of the 6,074 wells summarized in figure 7 range from 1.2 in. to 12 in. The significance of figure 7 is the systematic increase in yield and yield per foot that coincides with an increase in depth and diameter. To better define the narure of the interactions that are indicated in figure 7, least-squares regression analysis was employed. Yield and yield per foot of well depth were treated as dependent variables to be explained in terms of well depth and well diameter with the additional factor of topographic setting to be considered. Including depth and- diameter and interaction terms based on depth and diameter, a total of 20 potential variables were tested in model combinations containing from two to six variables in any one model. The models finally identified as having the best properties and best predictive capabilities contained three variables. Models containing additional variables were only increasingly complex without offering much more in pre- dictive capability. 'The variance in the model of yield versus depth and diameter was reduced by subsctting the data according to the three topographic groups identified earlier and recomputing the regression coefficients to produce three regression equations of the general fonn: yield = a -b (depth) + c (depth X diameter) -d (depth2 x diameter) where a. b. c, and d are regression coefficients. The regression equations and contour plots of the trend surfaces defined by these equations are shown in figures 8, 9. and 10. 'The contour plots are limited lo the range of known data. There are no small-diameter wells in the data set deeper than the no-data boundary. The deepest well in the data set is a 6-in. diameter well that is 1,301 ft deep. A number of larger diameter wells in the data set are nearly as deep. 'The shallowest well is 20 ft deep and 6 in. in diameter. Information contained in figun,s 8, 9, and 10 repre- sents several significant new findings regarding drilled wells in the crystalline rocks of the Piedmont and Blue Ridge. The surfaces shown in these illustrations represent the best average fit through yield data that has considerable Analysis Rei.tins Well Yield to Conslnlction and Sitil'l!, Piedmont-Blue Ridge Provinces, North Carolina A 17 10 8 6 4 154 LINE OF EQUAL WELL YIELD-- Intervals 18 and 19 gallons per minute Analyala based on data from 806 wells 2 / I I YIELD •11.1570 -0.11337 (DEPTH)+0.04190 (DEPTH· DIAMETER) -2.09739• 10' (DEPTH 2• DIAMETER) 0 '----'-----'----'---'---'----'----'-----'---'--'----'-----L..---' 0 200 400 600 800 1,000 1,200 TOTAL WELL DEPTH, IN FEET Figure 8. Contour plot of trend surface showing relation between well yield, total well depth, and well diameter for wells that are located in draws and valleys. variation at any given point. That is, for a point on any of the three contour plots there may be several wells of the same depth and diameter, all having different yields. This is important in interpreting the significance of the axes of the yield surfaces and why the average yield for wells of a given diameter decreases to the right of the yield-surface axes. Take for example, a point on the surface of figure 9 (wells on slopes and flats) representing a well depth of 525 ft and a diameter of6 in. The predicted average yield at this point, which also is on the yield-surface axis, is 32 gal/min. If a 6-in. well were drilled to this depth and had no water, two things could be done: stop or drill deeper. If drilling were stopped, that zero yield would be averaged with the yields of all other 6-in., 525-ft wells, which would average about 32 gal/min. If the well is drilled deeper and finally obtains water, the yield of that well averaged with other wells of the same depth will be less than at the yield-surface axis. Thus, for a given diameter well, the yield-surface axis represents the depth at which the maximum average yield will be obtained. Beyond the depth indicated by the axis, the chances of obtaining significant amounts, or additionala- mounts, of water decrease rapidly. This is perhaps better illustrated by figure 11 which is in effect a cross section of figures 8, 9, and 10. The figure is for a narrow range of well diameters. average 6 in., and shows the average yield and yield per foot for wells in intervals of well depth. The large data base of wells having diameters between 5.5 and 6.5 in. provides a sufficient number of wells in each depth interval to give a consistent picture and reduce scatter. A maximum average yield is reached in the interval between 500 and 550 ft (fig. 11 ). which is the approximate location of the yield-surface axes for 6-in. wells in figures 8, 9, and 10. The likelihood of A18 Ground-Water Resources of the Piedmont-Blue Ridge Provinces of North Carolina rn w 10 ~ 8 ~ i!:; ffi 1-w ~ :;!; C ..J 6 uJ 4 ~ 7 113 LOIE OF EQUAL WELL YIELD-- lntervala 15 and 1 e gallons par minute Analyala baaed on data from 2,681 wells 2 YIELD•8.1948 -0.11329 (DEPTH)+0,03494 (DEPTH-DIAMETER) -1.50931 • 1()5 (DEPTH~ DIAMETER) 0 '----'---'----'----''----'----'--...1....----'--...I....----'----'-----'------' 0 200 400 600 800 1.000 1.200 TOTAL WELL DEPTH, IN FEET Figure 9. Con lour plot of trend surface showing relation between well yield, total well depth, and well diameter for wells that are located on slopes and flats. obtaining significant additional quantities of water from 6-in. diameter wells decreases rapidly below depths of 550 ft. However. the increase in yield with increasing depth (up to the optimum depth) docs not occur in proportion to depth but actually decreases as the ratio to depth. By subsetting the well data by topographic groups, the regression analysis has resulted in three graphs (figures 8. 9, and 10) that at any well depth and diameter retain the relative magnitudes of yields identified in table 4. At any position on lhe graphs, the average yield for wells in valleys and draws is nearly three times the yield for wells on hills and ridges. The yield for wells on slopes and flats falls in between. Al1hough then: are differences in yield, the yield-surface axes of the three contour plots are nearly coincident, suggesting that topography may have little effect on the depth at which the maximum average yield is attained. 'The real significance lies in the position and shape of the yield-surface axes, which indicate that (I) well yield increases with depth to a much greater depth than previ- ously thought and (2) well yield increases dramatically as well diameter increases. The curvature of the yield-surface axes shows that depth is still a limiting factor, especially at depths greater than 500 to 600 ft as the axes of the yield surfaces rapidly curve away from the depth axes. However, the maximum average yield for I 2-in. wells is reached between 700 and 800 ft. This is much deeper than previ- ously thought. Cressler and others (1983) recently described similar large-diameter, deep, high-yield wells from the Piedmont of Georgia. Even the depth at which 6-in. wells reach their maximum average yield (about 500 ft) is 200 ft deeper than is usually recognized in the literature (LeGrand, 1967; Snipes and others, 1983). Analytis Relating Well Yield lo Construction and Siting, Piedmonl-lllue Ridge Provinces, North Carolina A 19 2 I I / / / LINE OF EQUAL WELL YIELD-- Interval ■ e ~nd 7 gallons per minute Analyala baaed on data from 1,880 wella YIELD• 8,882 -0.05303(DEPTH+O.O 1588)(DEPTH • DIAMETER) -7. 1124 7•106 (DEPTH\ DIAMETER) 0 L.___.L,__....1,__-1-_.-1. __ .___.L,_ _ _._ _ _._ _ __,_ __ .__ _ _._ _ __._ _ __, 0 200 400 600 800 1.000 1200 TOTAL WELL DEPTH, IN FEET Figure 10. Contour plot of trend surface showing relation between well yield, total well depth, and well diameter for wells that are located on hills and ridges. Although the regression analysis indicates that aver- age well yields continue to increase at greater depths than previously thought, perhaps the most interesting finding is the dramatic increase in average yield with an increase in well diameter. The effectiveness of increasing well diame- ter as opposed to drilling to greater depths is illustrated in figure 12, which is the result of a regression analysis of yield per foot versus well depth and diameter. The equation was derived in the same manner described earlier for the yield versus well depth and diameter relations. For a well of a given diameter, the yield per foot of hole is inversely proportional to the depth of the well, indicating that the amount of additional water obtained by drilling deeper is continuously decreasing. For wells of the same depth, however, increases in diameter are directly proportional to increases in yield per foot of well. Well Yields by Hydrogeologic Unit Well yields were matched to rock types to determine the relative yields of the different hydrogeologic units. The yield data were simultaneously sorted by topographic group to compare the relative importance of hydrogeologic unit versus topography as a consideration in selecting sites for wells. The results of these computations to compare yield, hydrogeologic unit, and topography are presented in table 7. Because yield is strongly influenced by well depth and diameter, which can lead to cultural bias favoring one hydrogeologic unit over another, a series of calculations was performed to remove the variation in well yield attributed to differences in depth and diameter. By using the equations (figs. 8, 9, and 10) relating well yield to depth and diameter for the three major topographic groups, the well yields were adjusted to an average 154-ft depth and A20 Cround-Water Resources of the Piedmont-Blue Ridge Provinces of North Carolina 50 ,----....----.-----.------,----~--~----. w I-::> z :i a: w CL en z 0 ..I ..I < C, ;!; c5 ..I !!:! >- w C, < a: w ~ 40 30 20 10 0 200 400 "' .... 0 ci "' 600 Analysis based on data from 4,298 wells 800 1.000 1.200 1.400 TOTAL WELL DEPTH, IN FEET Figure 11. Variation of average yield and yield per foot of well depth with depth for wells having diameters between 5.5 and 6.5 in. 6-in. diameter, the average of all wells in the data set. Because the influence of topography on well sites in the Coastal Plain is uncertain. the yields of wells in the Coastal Plain category were adjusted by using a regression equation that was computed for the entire data set and disregards topographic setting. It is nearly the same as the equation for wells on slopes and flats. The hydrogeologic units !IT (intermediate composition igneous intrusives), MBL (mar- ble), and SLT (slate) each had fewer than 15 observations having the necessary data (depth, diameter, yield, topogra- phy) to adjust the yields. Statistics for these hydrogeologic units, therefore, are not given, although the yields were included in the summary statistics. A regression of adjusted yields on hydrogeologic units is shown in figure 13. The average yields range from 23.6 gal/min for SCH (schist) to 11.6 gal/min for TRI (sedimentary rocks of Triassic age). The average difference in yield between adjacent hydrogeologic units in the regres- sion is only 0.6 gal/min. However, owing to the effect of the large number of wells in the analysis, the hydrogeologic unit can be used as a statistically reliable estimator (0.99 confidence level) of average well yield. Analysis-of•variance tests were also used to deter- mine whether any hydrogeologic units were significantly different from other hydrogeologic units in terms of yield. Because the average yields of all hydrogeologic units are not very different and the range of yields within units is very large, only those units toward opposite ends of the distri- bution are statistically different (0.95 confidence level) as indicated by the inequalities in figure 13. Three groups of hydrogeologic units stand out in figure 13. The metavolcanic units and ARO (argillite) form a group at the low end of the graph with only TRI (sedimentary rocks of Triassic age) having a lower average yield. Midway in the range of yields are the igneous units. At average or slightly above average yields are the metaig- neous units and QTZ (quartzite). The Piedmont crystalline Analysis Relating Well Yield to Construction and Siting, Piedmont-Blue Ridge Provinces, North Carolina A21 10 2 0.02 --- LINE OF EOUAL WELL YIELD-- Interval la varlable, gallona per minute per foot Analyala baaed on data from 5,077 wells YIELD PER FOOT• 0.002443 -S.8582 ><105 (DEPTH) -t-0.012997 (DIAMETER)+1.2S813 (DIAMETER/DEPTH) 0 '---.L--.L--..l.--..L.---'----'----'----L---L---L--....l__....[_ _ _J 0 200 400 600 800 1.000 1.200 TOTAL WELL DEPTH, IN FEET Figure 12. Contour plot of trend surface showing relation between yield per foot of well depth, total well depth, and well diameter. rocks underlying the Coastal Plain have the second highest average yield regardless of differences in crystalline rock composition. The high yield of these wells is attributed to the greater saturated thickness of overburden, which at an average 47. 7 ft is 1.8 times thicker than the 26.8-ft average for the rest of the study area based on 2,391 observations. including wells for which topographic information was not available. Well Yields by Geologic Belts and Terranes Comparison of well yields from the various geologic belts and terranes generally reflects the average yield of the predominant hydrogeologic unit(s). The yield data that were used for this comparison also were corrected to an average 154-ft depth and 6-in. diameter. A regression analysis of well yields in the various belts is shown in figure 14. The average difference in yield between belts is 0.9 gal/min. Average yield varies from a low of about 11.5 gaVmin for the Smith River allochthon (SR) and Triassic basins (TR) to a high of about 25.5 gal/min for the Murphy (MU). Analysis of variance tests found that the average yield of belts at the upper and lower ends of the data are signifi- cantly different. The inequalities significant at the 0.95 confidence level are also shown in figure 14. The belts having the highest yields, the Murphy (MU), Blue Ridge (BR), Chauga (CA), and Inner Piedmont (IP). are dominated by medium to high rank metasedimen- tary rocks, mafic gneisses, schists, and quartzitcs, and they include smaller ateas of metaigneous rocks, all of which have above average yields. The Charlotte belt (CH), which is characterized by igneous rocks intruded into country rocks of metavolcanic and metaigneous origin (Fullagar, 1971 ). and the Carolina slate belt (CS), which is dominated A22 Crouncf.Water Resources of the Piedmont•Blue Ridge Provinces of North Carolina Table 7. Relation of well yields to hydrogeologic unit and topography [Yield data arc adjusted to account for differences in yield due to d.iffcrecces in well depth and diameter. 'The average well is 6 in. in diameter and I S4 ft deep. The hydrogcologic unit!. are described in table I; gaVrnin, gallons per minute} Mean yield by toposraphic group Yield of all wells Hydrogeologic {gaVmin) ( al/min) Number of unit Draws and Slopes and Hills and Average First Median Third Ninth wells valler:s flats ridges guartile guartile dectile ARO ............ 26.8 16.3 12.5 14.6 7.0 11.5 17.0 27.0 319 CPL' ........... 21.7 9.1 14.5 21.8 37.2 419 GNF ............ 28.3 16.6 11.5 17.4 6.4 12.3 22.3 35.9 741 GNM ........... 33.5 19.6 12.3 19.9 6.5 12.5 23.4 40.7 1.129 IF!.. ............ 24.8 17.8 12.6 17.7 8.1 15.8 23.4 34.4 412 111' ............. 7 IMl2 ............ 24.4 12.1 17.8 4.7 14.0 19.9 44.0 29 MBL' ........... 3 MIF ............ 27.6 20.5 12.4 19.1 7.8 14.0 22.5 35.6 791 MIi ............. 22.1 20.6 ~13)'.;ti 18.4 8.8 16.0 23.3 36.2 284 MIM ............ 26.0 21.6 12:s 19.7 10.2 16.9 28.9 36.7 85 MVE'·· 16.6 11.9 16.9 7.5 11.8 16.0 25.0 95 MVF, ........... 19.0 15.1 9.5 13.0 6.2 11.2 17.8 25.9 280 MVI 17.1 15.5 16.8 9.2 13.4 23.6 35.2 43 MVM;:::::::::: 17.8 7.2 11.9 4.6 7.9 17.4 24.6 63 MVU ........... 27.1 23.4 10.9 20.2 8.1 14.8 24.5 41.2 141 PHL ............ 22.9 21.5 13.6 20.3 9.9 14.5 25.4 44.2 127 QTZ' ........... 20.6 16.8 18.6 4.8 15.2 29.4 46.5 65 SCH ............ 43.3 20.8 11.4 33.6 7.8 15.3 27.5 43.6 199 SL'f" ........... 2 TRI. ............ 19.0 12.2 8.5 11.6 4.7 9.0 14.5 25.5 269 All types 28.7 19.0 11.8 18.2 7.9 13.1 22.0 35.5 5,503 1Topography of bed.rock surface cannot be determined. Influence of topography on well yield in CPL area is unknown. zStatistics for categories having lc!is than 15 observations are not given. by metavolcanic rocks (Butler and Ragland, 1969), both are belts having low average yields. The areas containing sedimentary rocks, the Triassic basins (TR) and the western edge of the Coasrnl Plain (CP). are far apart in average yield, with the Triassic basins having the next-to-lowest yield and the Coastal Plain the third highest. SUMMARY AND CONCLUSIONS A statistical analysis was made of data from more than 6,200 wells drilled into the crystalline rocks of the Blue Ridge, the Piedmont, and the western edge of the Coastal Plain where crystalline rocks underlie sediments at shallow depths. This analysis was made to identify factors that are associated with high-yield wells. The data were classified according to geologic belts, hydrogeologic units composed of similar rock types.topographic selling, total and saturated thickness of regolith, water level, casing depth, yield, total depth, well diameter, and waler use. Six topographic seltings were combined into three groups based on well yields: hills and ridges, slopes and flats, and draws and valleys. Wells on hills and ridges had the lowest yields (averaging about 10 gal/min): wells in draws and valleys, the greatest (averaging about 30 gaVmin). Regolith thickness was about the same regardless of topographic group. but saturated thickness was least (about 19 ft) under hills and ridges and greatest (about 34 ft) under draws and valleys. Average yields in the geologic belts and hydrogeologic units ranged from about 11 to 25 gaVmin. There was considerable scatter in yields in all geologic belts and hydrogeologic units. Of 14 geologic belts, 10 were statistically different on the basis of well yield, as were 8 of 21 hydrogeologic units. About 70 percent of the wells were drilled for domestic use and, on the average, yielded about 11 gal/min; 80 percent of these wells were located on hills and ridges. The 30 percent of the wells drilled for public supply and commercial-industrial supply yielded about 30 gal/min on the average; about 50 percent of these wells were located in draws and valleys. The domestic wells had an average depth of about 125 ft; the public-supply and commercial- industrial wells, about 225 ft. Fewer than 2 percent of the domestic wells were 8 in. in diameter or larger. whereas nearly 25 percent of the public-supply and commercial- industrial wells were 8 in. or larger. Selecting the most favorable hydrogeologic unit or geologic belt alone can improve the chance of increasing the yield of the average 6-in. diameter, 154-ft deep well from Analysis Relating Well Yield to Construction and Siting, Piedmont~Blue Ridge Provinces, North Carolina A23 I:; .. C'I i! " :I: i .. I a 0 -1 ::!! a. ~ ~ i .. ,I ~ .. a 1· 2. z i ,.. n w a = ~ w I-:::, z ::E a: w 11. en z 0 ..J ..J < C, z c:i ..J !:!:! > w C, < a: w ~ 25,--,---,.----,---,----,-----,------r----r-----r--~--,---,---,---,----,----,---,......--,--~ 24 22 20 18 16 14 12 gallons per minute lnequalltles below were Identified by analysls ol variance at the 95 percent confidence level SCH ;t:ARG, MVF, MVM, TRI CPL, GNM ;i:ARG, MVF, TRI MIF:;t:MVF, TRI Figure does not Include the hydrogeologlc unlta marble (MBL), Igneous intrusive intermediate (Ill) and slate (SL T) owing to lack ol data The hydrogeologic units are described in table 1 • • • • AVERAGE YIELD•-0.588 (HYDROGEOLOGIC UNIT)+23.14 Analysis based on data from 5,489 wells • IGNEOUS • • METAIGNEOUS ROCKS ROCKS METAVOLCANIC ROCKS ~ CPL PHL MVU GNM MIM MIF QTZ MIi IMI IFI GNF MVE MVI ARG MVF MVM TRI 10'--.J__.J__.J__J_ _ _L.. _ _L.. _ _L.._...J...__....J...._....J...._....J.... _ _,_ _ _._ _ _._ _ _i _ _i _ _i _ __J _ __J SCH 6 7 8 9 10 11 12 13 14 15 16 17 18 1 2 3 4 5 HYDROGEOLOGIC UNIT figure 13. Average yield of wells of average construction in the hydrogeologic units of the Piedmont and Blue Ridge provinces of North Carolina. ~ ii: .; I f i ~ ~ f .. c ~ f a l w ~ ~ ::!: 0:: w a. rn z g ..J < C, ~ ci ..J !Y >-w C, < 0:: w ~ 26i-----.--,-----,--,-----,--,-------r--.------r--,--------r--,----~--~-~ Inequalities below were identified by analysis of variance at the 95 percent confidence level 24 TR, Ml ;t BR, CP, CA. IP, SA, RA, CH CS ;t BR, CP, IP 22 The geologic belts and terranes are described • In table 2 and shown in figure 8 20 • • 18 MEAN 18.24 gallons per minute • 16 AVERAGE YIELD•25.23 -0.900 (BELn 14 Analysis based on data from 5,498 wells • 12 • • MU BR CP CA IP SA GH RA CH KM cs Ml TR SR 10 1 2 3 4 5 6 7 8 9 10 11 12 13 14 GEOLOGIC BELT OR TERRANE ! Figure 14. Average yield of wells of average construction in the geologic belts and terranes of the Piedmont and Blue Ridge provinces of North Carolina. "' i about 11 to 12 gal/min to about 24 to 25 gal/min. about a twofold increase. Considering topography alone, the average well on hills and ridges can be expected to average less than 12 gal/min, whereas wells in draws and valleys can be expected to average about 29 gaVmin, an increase of 2.4 times. When the factors of hydrogeologic unit or geologic belt are considered in combination with topo- graphic setting, the range in yields is even greater. Wells in draws and valleys in the most productive units average five times more yield than wells on hills and ridges in the least productive units. The statistical analysis supported some concepts and criteria for well-site selection, such as the siting of a well with regard to topography. More importantly, however, the analysis indicates that some previously held concepts may be in error. First and foremost is the generally held concept that the crystalline rocks yield only small amounts of water to wells. The analysis showed that this concept may be due to cultural bias. Most wells drilled in these rocks are small diameter. are located primarily on hills and ridges-the poorest possible sites for wells-and are drilled only to depths where sufficient water for a domestic supply is obtained. In the same theme, well diameter has not been considered to have much effect on yield-a large-diameter well was considered a storage tank. Statistical analysis shows, however. that for a given depth the yield of a well is directly proportional to the well diameter. The larger the diameter the greater the yield. Well construction in crystalline rocks has long been based on the concept of a well intersecting near vertical open fractures and joints that, because of lithostatic pres- sure, pinch out at depths of about 300 ft. As a result, the drilling of many wells has been arbilrarily stopped when the depth of 300 ft was reached. The average well, whether domestic or commercial-industrial. is not even that deep. The analysis indicates that very few wells have been drilled deep enough to test the full potential of the sites. For example, the average yield of 6-in. diameter wells located in draws or valleys reaches a maximum of about 45 gaVmin at depths of 500 to 525 ft; the average yield of 12-in. diameter wells located in draws or valleys reaches a maximum of about 150 gal/min at depths of 700 to 800 ft. Whatever the hydrogeologic unit or topographic loca- tion, the chances of obtaining high yields are enhanced by increasing the depth and diameter of the well to a much greater extent than previously thought. REFERENCES Bain, G.L., 1966, Geology und ground-water resources of the Durham area, Nonh Carolina: Nonh Carolina Department of Water Resources Ground-Water Buhetin 7, 147 p. Brown. P.M .. and Pwlcer, J.M., Ill, compiler.;, 1985, Geologic map of North Carolina: North Carolina Department of Natu- ral Resources and Community Development. 1 sheet. scale 1:500,000. Burt, E.R .. 1981. Geologic map of Region H. North Carolina: North Carolina Geological Swvey. Open-File Map no. NCGS 81-4, scale 1:125,000. Butler. J.R., and Ragland. P.C .. 1969, A petrochemical survey of plutonic intrusions in the Piedmont, southeastern Appala• chians. U.S.A.: Contributions to Mineralogy and Petrology. v. 24, p. 164-190. Carpenter, P.A., Ill, 1982, Geologic map of Region G, Nonh Carolina: Nonh carolina Department of Natural Resources and Community Development, Geological Survey Section. Regional Geology Series 2, scale 1:125,000. Cedarstrom. D.J., 1972, Evaluation of yields of wells in consol- idated rocks, Virginia to Maine: U.S. Geological Survey Water-Supply Paper 2021. 38 p. Conrad. S.G .. Carpenter, P.A., Ill. and Wilson, W.F .. 1975, Physiography, geology and mineral resources. in Clay, J. W ., Orr, D.M .. and Sruan. A.W., eds .. North Carolina atlas, portrait of a changing southern State: Chapel Hill. North Carolina. University of North Carolina Press, p. 112-127. Cressler. C.W., Thurmond. C.J., and Hester. W.G .. 1983. Ground water in the greater Atlanta region: Georgia Depart• ment of Natural Resources Environmental Protection Divi- sion, Georgia Geologic Survey Information Circular 63. 144 p. Daniel. C.C., III. and Sharpless, N.B., 1983. Ground-water supply potential and procedures for well-site selection in the upper Cape Fear River basin. North Carolina: North Carolina Department of Natural Resource$ and Community Develop- ment, 73 p. Dodson, C.L.. and Laney. R.L., 1968. Geology and ground-water resources of the Murphy area, Nonh Carolina: North Carolina Department of Water and Air Resources Ground-Water Bulletin 13. I 13 p. Eder. B.K .. Davis. J.M .. and Robinson, P.J., 1983. Variations in momhly precipitation over North Carolina: University of North Carolina Water Resources Research Institute Report no. 185. 50 p. fapenshadc, G.H., Rankin. D.W .. Shaw, K.W .. and Neuman, R.B., 1975. Geologic map of the east half of the Winston- Salem quadrangle. Nonh Carolina-Virginia: U.S. Geological Survey Miscellaneous Investigations Series Map 1-709-B. scale 1:250,000. Fenneman. N.M., 1938, Physiography of Eastern United States: New York, McGraw-Hill, 714 p. Floyd. E.O., 1965. Geology and ground-water resources of the Monroe area. North Carolina: North Carolina Department of Water Resources Ground-Water Bulletin 5, 109 p. Fullagar. P.O., 1971. Age and origin of plutonic intru5ions in the Piedmont of the 50Utheastcm Appalachians: Geological Soci- ety of America Bulletin, v. 82. p. 2845-2862. Goldsmith. Richard. Milton. D.J .. and Horton. J.W .. Jr .. 1982, Simplified preliminary geologic map of the Charlotte IO x 2° quadrangle, North Carolina and South Carolina: U.S. Geo- logical Survey Open-File Report 82-56. scale I :250.000. Hadley, J.B .. and Nelson. A.E., 1971. Geologic map of the Knoxville quadrangle, North Carolina. Tennessee, and South A26 Ground-Water Resources of the Piedmont-Blue Ridge Provinces of North Carolina Carolina: U.S. Geological Survey Miscellaneous Investiga- tions Series Map 1-654, scale 1:250,000. Heath, M.S .. and Coffield, H.J .• Ill, 1970, Cases and materials on ground water law-with particular reference to the law of North Carolina: Chapel Hill, North Carolina, University of North Carolina Instilute of Government, 72 p. Heath, R.C., 1980. Basic elements of ground-water hydrology with reference to conditions in North Carolina: U.S. Geolog- ical Survey Water-Resources Investigations Open-File Report 80-44. 86 p. LeGrand. H.E., 1954, Geology and ground water in the Statesville area. Nonh Carolina: North Carolina Department of Conser- vation and Development Bulletin 68, 68 p. --1967, Ground water of the Piedmont and Blue Ridge provinces in the southeastern States: U.S. Geological Survey Circular 538. 11 p. LeGrand, H,E., and Mundorlf, M.J .. 1952. Geology and ground water in the Charlotte area. North Carolina: North Carolina Department of Conservation and Development Bulletin 63, 88 p, Marsh, Q.T .. and Laney, R.L., 1966, Reconnaissance of the ground-water resources in the Waynesville area. North Caro- lina: North Carolina Department of Water Resources Ground- Water Bulletin 8. 131 p. May, V.J., and Thomas, J.D .. 1968. Geology and ground-water resources in the Raleigh area, North Carolina: North Carolina Department of Wat~r and Air Resources Ground-Water Bulletin 15, 135 p. McDaniel, R.D .. 1980, Geologic map of Region K, Nonh Carolina: North Carolina Geological Survey. Open-File Map no. NCGS 80-2. scale 1:125,000. Mundorlf. M.J ., I 946, Ground water in the Halifax area, Nonh Carolina: North Carolina Department of Conser.·ation and Development Bulletin 51, 76 p. --1948, Geology and ground water in the Greensboro area. North Carolina: North Carolina Department of Conservation and Development Bulletin 55. 108 p. Peace, R.R .. Jr., and Link, D.R., 1971, Geology and ground- water resources of northwestern North Carolina: North Caro• lina Department of Water and Air Resources Ground-Water Bulletin 19. 135 p. Pusey, R.D., I 960. Geology and ground water in the Goldsboro area, Nono Carolin,: Nonh Carolina Depanment of Water Resources Ground-Water Bulletin 2. 77 p. Ragland, P.C .. Hatcher, R.D .. Jr., and Whittington, David, 1983, Juxtaposed Mesozoic diabase dike sets from the Carolinas: A preliminary assessment: Geology, July 1983, v, 11, p. 394-399, Rankin. D.W., Espenshade. G.H .. and Neuman, R.B .. 1972, Geologic map of the west half of the Winston-Salem quad- nmgle, North Carolina, Virginia. and Tennessee: U.S. Geo• logical Survey Miscellaneous Investigations Series Map 1-709-A, scale 1:125,000. Reinemund. J.A., 1955, Geology of the Deep River coal field, Nonh Carolina: U.S. Geological Survey Professional Paper no. 246. 159 p. SAS Institute Inc., 1982a. SAS user's guide: basics. 1982 edition: Cary, North Carolina. SAS Institute, Inc., 923 p. --1982b, SAS user's guide: statistics. 1982 edition: Cary. Nonh Carolina, SAS Institute. Inc .. 584 p. Schipf, R.G., 1961, Geology and ground-water resources of the Fayetteville area: Nonh Carolina Department of Water Resources Ground-Water Bulletin 3, 99 p, Seiders, V,M .. 1981, Geologic map of the Asheboro, Nonh Carolina, and adjacent areas: U.S. Geological Survey Mis• cellancous Investigations Series Map 1-1314, scale I :62,500, Snipes. D.S .. Padgett, G.S .. Hughes, W.S., and Springston. G.E., 1983, Ground•water quantity and quality in fracture zones in Abbeville County. South Carolina: Clemson. South Carolina, Clemson University Water Resources Research Institute Technical Repon no. l02, 54 p. Steel. R.G.D .. and Torrie, J,H .. 1960. Principles and procedures of statistics, with special reference to the biological sciences: New York, McGraw-Hill, 481 p. Stromquist. A.A .. Choquette. P.W .. and Sundelius, H.W .. 1971. Geologic map of the Denton quadrangle, central Nonh Carolina: U.S. Geological Survey Geologic Quadrangle Map GQ-872, scale I :62,500. Stromquist, A.A .. and Sundelius, H.W .. 1975. Interpretive geo- logic map of the bedrock showing radioactivity, and aero- magnetic map of the Salisbury, Southmont, Rockwell. and Gold Hill quadrangles, Rowan and Davidson Counties. North Carolina: U.S. Geological Survey Miscellaneous Investiga- tions Series Map 1---888 (sheet 1 of 2). scale I :48,000. Stuckey. J.L .. 1958, Geologic map of Nonh Carolina: Raleigh, North Carolina. North Carolina Department of Conservation and Development, Division of Mineral Resources. scale 1:500,000. Sumsion, C.T .. and Laney, R.L .. 1967, Geology and ground- water resources of the Morganton area: North Carolina Department of Water Resources Ground-Water Bulletin 12, 119 p, Trapp. Henry, Jr .. 1970, Geology and ground-water resources of the Asheville area, North Carolina: North Carolina Depart- ment of Water and Air Resources Ground-Water Bulletin 16, 127 p. Weigand. P.W., and Ragland, P.C., 1970, Geochemistry of Mesozoic dolerite dikes from eastern North America: Contri- butions to Minerology and Petrology, v. 29, p. 195-214. Wilson. W.F .. 1979. Geology of Wilson County, Nonh Carclina: Nonh Carolina Depanment of Natural Resources and Com- munity Development. Division of Land Resources. Geolog- ical Survey Section, Open-File Map NCGS 79-2, scale 1:125,000. --1981, Geology of Halifax County, Nonh Carolina: Nonh Carolina Geological Survey, Open-File Map NCGS 81-3, scale 1:125,000. Wilson. W.F .. Carpenter, P.A .. Ill, and Parker. J.M., III. 1981, Geologic map of Region J, North Carolina, in A guide for North Caro1ina mineral resource development and land use planning: North Carolina Department of Natural Resources and Community Development. Geological Survey Section, Regional Geology Series I, scale I: 125,000. Wilson. W.F .. and Spence, W.H., 1979, Geology of Nash County. North Carolina: North Carolina Department of Nat• ural Resources and Community Development, Division of Land Resources, Geologica1 Survey Section, Open-File Map NCGS 79-3, scale 1:125,000. Analysis Relating Well Yield to ConstNction and Siting. Piedmont-Blue Ridge Provinces, North Carolina A27 APPENDIXC II PROJECT NAME YAN I!!UCK It !!US DATE STARTED 3/6/91 COMPLETED 3/7/91 PAOE...LOF_II_ JOB NO. 36 COUNT¥ ROWAN CITY/TOWN CU:Vli:LAIID · STATE NC BORING/WELL lllf-1 TOTAL DEPTH 20,0 BORING LOCATION UP OIWllENT ELEV. 1nu, FT GEOLOGIST /ENGINEER BRIAN TEMPLE MONITORING EQUIPMENT USED OVA DRILLERS WELL DRIU.EB~ INC DRIWNG EQUIPMENT 8 112 RIO METHOD UOb!,Olt g)'.f;!! AUG!;I! . REMARKS: ALL MEASUREIIENTS ARE DELOY GROUND SURFACE . . . DEPTH TO BOTTOM OF SCREEN 12;0 FT TOP OF SCREEN 2.0 FT SCREEN SLOT SIZE .01 IN DEPTH TO BOTTOM OF SAND . 20.9 n TOP OF SAND UI E1 DEPTH TO TOP OF BENTONITE 119 [[ WELL MATERIAL USED PVC DIAMETER 2.00 !!! STICK UP 2.0[! WATER ZONES (DEPTH) B.O l'I' STATIC WATER LEVEL ft.2& l'I' REMARKS: ALL MEASUREMENTS ARE BELOW GROUND SURFACE DEPTH SAMPLE BLOW % OVA LITHOLOGY (FT) NO. COU T REC. SAMPLE DESCRIPTION READING 6" 6" 6'. (PPM) OIIASS FIRST 2 CII SUOIITLY IIICACEOUS -5 I I I 3 t.B7F'I IIIIITE, DARK GRBEII, BRO'IN 7 1101ST, CLAYEY, F1NE BANDY, SILT (6-Mn) WATER Ill 8 FT -10 I 2 3 6 1.10 n IIIIITE, DARK GREEII, BRO'IN 1101ST, CLAYEY, F111E/IIEDIUII SAIIDY ,3 SILT (to-11.11 rr) -15 2 4 8 2.20n IIETA-DIORITE 3 . l 1101ST, CLAnY, COARSI SANDY, SILT (16-18.6 FT) ~ 20 BOIUNO TERMINATED O 20,0 FT -25 IIOTE, ALL S00.S IOOIIB1111D IGH&:OUS TEXTURES AT Dl!!PTIIS Bl:LOY 4,0 FT '-. 30 -35 '-40 -45 SAMPLES SUBMmED FOR LABORATORY TESTING .QUATERRA, INC. I 2 (HONE) . ALEJOH,OREENSBORO,CHARLOTTE 3 NORffi CAROIJNA APPENDIXD Classification of Transmissivity Magnitude and Variation by Jiri Knlsny" Abstract Until now no objective claallcation of transmis.,ivity has bun introduced, in spite of the quantitative nature of transmissivity values and their ohl'ious importance for quantitative appraisals of aquifen or major ground-water systems. The usual subjective expres.sion of transmls.sivity, for example as "high" or "low," prevents the objective comparison of transmlssivlty values characterizing different areas and hydrogeological environments. A combined da'5iflcation of magni- tude and variation of transmissivity is proposed, with the intention to standardize the expression, comparison, and representation oftrall'lmlsslvlty. This classification also enables Its compact and unambisuous depiction in tables and maps. Introduction Transmissivity is an important hydraulic property of aquifers and water-bearing materials. In common with permeability, transmissivity affords a notion about the water-bearing characteristics of hydrogeological bodies. Transmissivity values enable us to estimate the possibility of ground-water abstraction, in the first approximation. Therefore, knowledge of transmissivity distribution helps us to draw important conclusions from hydrogeological stu- dies, and for this reason, prevailing transmissivity values arc often represented in hydrogeological maps. They provide a basis for future ground-water exploration, development, abstraction, and protection. Yet, in spite of the quantitative nature of transmissivity and its importance for quantitative appraisals, no objective classification of transmissivity has been introduoed. Quan- titative or semiquantitative terms describing transmi~ivity are often used, denominating different grades or classes as large, small, etc., but without strictly stating limils between them. This is the current case with hydrogeological maps, where the inexactly defined term "productivity" is some- times used (cf. e.g., !AH et al., 1983). Even iftransmissivityis exprtSSCd numerically, the verbal designation of numerical classes might involuntarily reflect the relation between hydrogeological conditions and water demand; in areas where yields of water wells are sufficient to cover limited water consumption, transmissivity may be designated as high; on the other hand, where well yields do not suffice for • Department of Hydrogealo~ and Engineering Geology, Faculty of Science, Charles University, Prague, Czech Republic, Received January 19921 revised July 1992, accepted Sep- tember 1992. Discussion open until September I, 1993. 230 large requirements, transmissivity might be designated as low. Such a subjective approach prevents the objective comparison of transmissivity at both local and regional scales, including values represented on hydrogeological maps. Just as with transmissivity magnitude, transmissivity variation affords important information on hydrogeological properties. In spite ofits usefulness, however, this character- istic is very seldom used in hydrogeological studies, and no objective basis for its comparison has been defined. Therefore, a combined classification of transmissivity magnitude and variation is proposed with the intention to standardize transmissivity expression, representation, and comparison in regional and local hydrogeological studies. This classification aiso enables its compact and unambigu- ous depiction in tables and r,naps. Background The statistical distribution and prevailing values of permeability and transmissivity strongly depend on the rela- tion of the magnitude of elements of rocks heterogeneity to the extent of the studied area (Rats, 1967). The idea was hydrogeologically interpreted by Kiraly (1975). Conse- quently, because of this "scale effect," representative values of hydraulic parameters are substantially influenced by methods used for their determination. Pumping tests arc the most frequent procedure used to determine transmissivity values that are often available as extensive data populations that may be used to characterize a hydrogeological envi- ronment. Wells are also the most common device for ground-water abstraction in general. Many times, however, various less reliable historical data are available in arehives. These data are not suitable for determining exact hydraulic parameters such as the transmissivity and hydraulic conduc- tivity; on the other hand, many of these data are good enough to contribute to a general idea about the regional Vol. 31, No. 2-GROUNO WATER-March-April 1993 Table I. Classllleatlon of Transmlsslvlty Magnitude Comparati~ regional pararntttrs approximaltly corresponding to the coefficient of transmissivlly Nonlogarithmic Coeflic~nt of Class of Designation of ----- tr01Umissivl1y tran.rmlsslvily transmisslvity ~ciflc capacity (m 2/d) magni1ude magnitude q in I/rm I Very high -1,000 JO II High 100 Ill lntenncdiate JO 0.1 IV Low 0.01 V Very low .0.1 0.001 VI Imperceptible distribution of transmissivity and/ or permeability values and could be treated statistically. Therefore, a "category" of comparative regional parameters was introduced (Jetel, 1964; Jetel and Krasny, 1968), as described below. This gives us a good chance of drawing regional conclusions, especially when using simple statistical procedures for data analysis. The classification of transmissivity magnitude and variation is proposed with the intention of expressing, representing, and comparing available transmissivity data in a more objective manner. Transmlsslvlty Magnitude This classification of transmissivity magnitude was published by Krasny ( 1970) and has been modified several times afterwards (e.g. Knlsny, 1986c) to reach the present form. On the basis of transmissivity studies both in Czechoslovakia and abroad, the range of values generally found is logarithmically divided into six classes from very high transmissivity (/ class: coefficient of transmissivity T more than 1,000 m2/d) to imperceptible transmissivity ( VI class: T less than 0.1 m2/d) as shown in Table I. In the table, the range of different classes is expressed in SI units and in approximately corresponding comparative regional parame- ters (i.e., specific capacity q and index of transmissivity Y) as well. In Figure I, English units are also shown for compari- son. The divisions between classes in English units follow Very approximate Logarithmic expected discharge ----in 1/s ofa Index Ground-water single well at y supply potential 5 mdrawdown Withdrawals of great regional importance >SO 7.0 Withdrawals of lesser regional importance 5 -50 6.0 Withdrawals for local water supply (small communities, plants, etc.) 0.5 -5 5.0 Smaller withdrawals for local water supply (private consumption, etc.) 0.05 -0.5 4.0 Withdrawals for local water supply with limited consumption 0.005-0.05 3.0 Sources for locaJ water supply arc difficult (if possible) to ensure <0.005 more or less the orders of magnitude of SI units only by chance, as for example, in the case of transmissivity expressed in m2/d and ft2/d. The index of transmissivity, Y, has been introduced as a comparative regional parameter, a logarithmic tramforma- tion of specific capacity (Jetel and Kr6sny, 1968). It has been used advantageously since the distributions of the majority of sample populations of transmissivity values are log- arithmic-normal; when using a logarithmic parameter (i.e., the index Y, for example), the statistical distribution changes to a simpler normal one. To calculate the index Y =log(l06 q), specific capacity q has to be expressed in 1/s m. Approximate estimates of probable ground-water yield to wells are appended to all classes in Table I. Thus, results of individual wells can be evaluated in the first approximation, but, more significantly, extended areas may be assessed in a preliminary way aocording to prevailing transmissivity values with respect to this important hydro- geological issue (Table 1). Many hydrogcological media are characterized by so- called chaotic heterogeneity; these are media where perme- ability (transmissivity) distribution is a random function without any apparent regional tendency, cf. Borevskiy ct al. ( 1979). Where there are sufficient available data to be pro- cessed statistically, sample populations delimited by differ- 231 ent rock types, areas, hydrogeological positions, etc. can be treated to determine the arithmetic mean x and the standard deviation of sample s of each of them. As the factor by which to determine the class of transmissivity magnitude for each population, the interval x ± s was chosen; if 70 or more percent of the interval belongs to one class, prevailing transmissivity is designated by the name or symbol of this class; if 30 to 70 percent of that interval belongs to either of the two (exceptionally three) adjacent classes, names (sym- bols) of both (all) classes are concentrated to designate the prevailing transmissivity, in order of the magnitude of their participation. In the case where 30 to 10 percent belongs to a class, its name (symbol) is enclosed in parentheses. No name (symbol) is used with a percentage less than 10. Some exam- ples of the classification procedures are given later. Trans- missivity values expressed by index of transmissivity Y or by coefficient of transmissivity or specific capacity, the last two in a logarithmic fonn, should be used when treating the samples statistically and determining the interval i ± s. Examples of graphical statistical treatment of different sample populations of transmissivity values are shown in Figure 1, where probability paper for plotting cumulative relative frequencies of samples was used. Similar representa- tion was used by Walton (1962, in Davis and De Wiest, 1966). More about the application of probability paper can be found in Spiegel (1972). Values of transmissivity outside the interval x ± s are considered anomalies, positive and negative (Figure I). Both are of practical importance: the positive anomalies (the interval between i + s and x + 2s, designated as + A in Figure I) indicate the zones of better possibilities for ground-water abstraction compared with the area of prevail- ing transmissivity (hydrogeological background). The nega- tive ones(values betweeni -sand x-2s, i.e. -A)showthc zones to be avoided for water-supply purposes or where, on the other hand, there is generally less danger of ground- water contamination, and where, consequently, location of landfills may be considered. The extreme anomalies, posi- tive(++ A) and negative ones (--A) can be found outside the interval i ± 2s (Figure I). In cases where data cannot be treated statistically, the class of transmissivity magnitude is estimated directly on the basis of prevailing values. Variation In Transmlaslvlty No homogeneous environment exists under natural conditions; all environments are heterogeneous to a differ~ ent extent often depending on the extent of the studied area CLASSES OF TRANSMISSIVITY MAGNITUDE VERY LOW LOW INTERMEOIATE HIGH VERY HIGH ++A V IV 111 II ¼ ~ ... ,.,., " +A .. --){+S "' "' ., +I ., ll IX " .. \t~ +a+. .. l!: ~ X/X " 701 ~ ~ e • • "' " N ..., '5 00 ,,; <D O> 1! ,0 8 I ~ ~ -'-X-s IS s ' 10 -A i ±s s X-25 -• I ll 13 Index y --A :i: 15 I .0 I ,s 5.0 ss 6.0 65 7.0 7.S 8. 00!)1~ .. ,m, ,nb;-~· I ,0/';I.~, ~°'' I -~·r1t bi I I i'J';,o (~ I 1~•11.lf(l/sml q I I 1 I i911 I 1 , hii , I I ill I I (gpd/ft l 1 tO SO :200 SOQ1 2 1'-2 15z S. 01 02 , : 0 Ill 1 ' 1 1 (m21dl , "' s 1 !frldl T Fl&, t. Cumulative relative frequencies of different samples of transmlsshrity values and their classification according to transmlsslvlty m11111ltude and variation, wbere q = 1pecillc capacity In 1/s m and in IJICl/ft; T = coefficient of transmlssivlly in m'/d and in lt'/d; i = arithmetic mean; s::::: standard deviation; and ++A1 +A, -A, --A~ fields of positive and neaative anomalies (extreme anomalies), respectively, and lndlvldual anomalous water wells (outdde tbe lntenal i ± s or prevailing tnnsmissivlty values = hydrogeoloalcal background). 232 Table 2. Claulflcatlon or TrammlssMty Variation Standard deviation of rransmwlvily index y• Class of .O.signation of Hydrogeological environment from the point of view of its hydraulic heterogeneity 0 transmissivity trarumissiviry wuiation variation a Insignificant Homogeneous --0.2--------------------------------- b Small Slightly heterogeneous --0.4--------------------------------- C Moderate Fairly heterogeneous --0.6--------------------------------- d Large Considerably heterogeneous --0.8------------------------------- e Vcry heterogeneous Very large --1.0------------------'-'---------------- f Extremely large Extremely heterogeneous • Or logarithmic transformation of any parameter cxpre.<ising transml~ivity. •• Usable especially for permeability evaluation but also when evaluating transmi~ivity. and the method used for hydraulic parameters assessment (scale effect). Variation in transmissivity can be expressed by using the standard deviation of the sample population of trans- missivity values. Six classes denominated a to/ are distin- guished, with all classes having the same range: 0.2 of the standard deviation of the logarithm of transmissivity. These classes are based on the results ofstudies in different hydro- geological media (Table 2). It is well-known that insignifi- cant or small variation (classes a, b) is typical of samples representing media having intergranular porosity ( e.g., well- sorted fluvial deposits) while large or very large variation of transmissivity (classes d, e) are characteristic of media with dominant fissure porosity. Some examples are shown in Figure 1: the steeper the slope of the line representing the sample, the smaller the transrnissivity variation of the sample. Moreover, knowledge of transmissivity variation and of the interval i ± s enables us to express a range of probable transmissivity values and/or intervals representing anoma- lies, and makes possible the prediction-on the basis of previous data--0f future water well results, assuming that wells are drilled in the same area under the same conditions as those used for the statistical analysis. Specific intervals of Y, q, or T values can be predicted on the respective axes, as shown in Figure I. The smaller the standard deviation (i.e., the narrower the interval i ± s), the more reliable is the prediction. The classification of variation (Table 2) may, of course, be used for the evaluation of permeability (hydraulic con- ductivity) as well, if this parameter is expressed in a loga- rithmic form. Variation of permeability and transmissivity expresses a character of hydrogeological environment from the point of view of its hydraulic heterogeneity. In the case of transmissivity it can also reflect changes in aquifer thickness. Cl888lllcatlon of Transmlsslvlty The classification of transmissivity magnitude ( six classes/-VI) combined with the classification of transmissiv- ity variation (six classes a-/) can be used for a quick and very simple quantitative characterization of different hydrogeo- logical environments and for their comparison. Jn addition, knowledge of transmissivity variation and of the interval i ± s enables us to predict future well yields and to indicate the hydraulic character of a hydrogeological environment. For example, an environment classified as la, i.e., very high transmissivity with insignificant variation, represents the optimum environment for ground-water development. Of course, this characteristic does not express natural ( or other) ground-water recharge. On the other hand, classification Via (imperceptible transmissivity with its insignificant varia- tion) indicates an environment almost without promise of ground-water abstraction possibilities, while Vie or Vlf (imperceptible transmissivity with very or extremely large variation) leaves some hope of discovering more permeable zones or areas where transmissivity may differ significantly from the mean values due to the extreme heterogeneity of the hydrogeological environment. The classification of transmissivity can be used in hydrogeological maps as reported in Krasny and L6pez (1989). It was used in the 1:500,000 hydrogeological map of the Czech republic and in a slightly modified version also in the new edition of hydrogeological maps of Czech territory at 1 :50,000 scale. In maps and in hydrogeological studies, three degrees may be distinguished of the density and reli- ability of available data: I. In areas with sufficient data to process statistically, the interval i ± s of prevailing transmissivity values (hydro- geological background) can be estimated; then both trans- missivity magnitude and variation can be expressed accord- ing to classifications in Tables I and 2, as mentioned above. 2. In areas where prevailing transmissivity is estimated on the basis of less numerous data so that no statistical treatment can be undertaken, only the class of transmissivity ., magnitude may be assessed and expressed, for example by symbols, as follows: /, Ill-IV. etc. I 3. In areas with few or no transmissivity data, the prevailing transmissivity is assessed by analogy or by con- sidering the geological environment (especially lithology and structure) of the area and expressed by a symbol in brackets, e.g. {VJ, [I-II}. 233 Table 3. Basic Data for the Examples Example: Sample: Geological environment: Values of the index Y: n: x: s: Index Y: i±s T (m2/d): Classes of transmissivity magnitude and variation of samples: I A Crystal/IN rocks (esp. gnt!isS) 3.32 3.43 3.76 3.80 3.82 4.37 4.40 4.43 4.56 4.78 4.85 S.15 S.18 13 4.30 0.62 3.68-4.92 0.45-7.9 IV(-V)d ] B Cretaceow sandstones 4.05 4.72 5.27 5.32 5.38 S.48 5.17 S.89 5.93 6.01 6.08 6.15 6.30 6.38 6.43 6.47 6.51 6.54 6.65 19 5.86 0.68 5.18~.54 14-330 lll-lld 3 C Quaternary fluvial deposits 6.39 6.51 6.53 6.76 6.80 6.82 6.83 6.84 6.90 6.90 6.92 7.05 12 6.77 0.19 6.58~.96 J(i()-860 Ila n = sample size; i = arithmetic mean of a sample; s = standard deviation of the sample; T = coefficient of transmissivity. All the values of the index of transmissivity Y (and accordingly those of the specific capacity, from which indices Y were derived) were determined as results of pumpina tests from water wells. Examples of TransmlBBlvtty ClaBBlllcatlon As examples of the clll&'lification procedure, and also to demonstrate actual differences in transrnissivity magnitude and variation, three real statistical samples oftransmissivity values from different hydrogeological environments within the Bohemian Massif in Czechoslovakia are presented. The basic data for the examples are given in Table 3. Forsimplic- ity, samples are characterized directly by the values of the index of transmissivity Y, i.e., by a logarithmic modification of specific capacity q, as stated above. The approximate relation between the index of transrnissivity Y, specific capacity q, and coefficient oftransmissivity T follows from Table I or Figure I. Example I. In an area formed by crystalline rocks (especially gneiss) in southern Bohernia(sample A in Figure I), the available values of index of transmissivity Y (fable 3) give the arithmetic mean x as 4.30, and the sample standard deviation s as 0.62. The classification interval i ± s covers the range 3.68-4.92, of which 74.2 percent belongs to the 234 class /Vof transmissivity magnitude and 25.8 percent to the class V. According to its standard deviation, the sample belongs to the class d of transmissivity variation. Then the designation of the sample is IV (-VJ d. i.e. low (to very low) transmissivity with large variation. The approximate conversion of the index Y values to the respective transmissivitycoefficient T may be performed using the equation T(in m2/d)= tov-u, X86,400(Jetel and Krllsny, 1968). Then the arithmetic mean of sample expressed as transmissivity Tis 1.89 m2/d, and the interval x ± s lies between 0.45 and 7.9 m2/d. Example 2. A sample of Y values from sandstones of the Cretaceous basin in northern Bohemia (sample B in Table 3 and Figure I) has the arithmetic mean 5.86 ( of index Y), or expressed as T = 69 m'/d. Standard deviation of samples =0.68; 60.3 percent of the interval x± s (5.18-6.54) belongs to the class Ill, 39. 7 percent to the class fl Accord· ing to the cla.~ification the sample may be designated a." Ill-lid, i.e., intennediate to high transmissivity with large variation. Approximately 68 percent of the sample (interval x ± s) can be expected to be between transmissivity values I 4 and 330 m2/d. Example 3. A sample ofY values from fluvial deposits of the River Labe in central Bohemia (sample C in Table 3 and Figure I) has an arithmetic mean x = 6. 77 (T = 650 m2/d), standard deviations =0.19, interval x ± s = 6.58-0.96 (f = 360-860 m2/d). One hundred percent of the interval belongs to the class II; according to the standard deviation, the class of transmissivity variation is a. The sample can be classified as //a, i.e., high transrnissivity with insignificant variation. Closing DlscuBBlon and Conclusions The proposed classification of transmissivity magni- tude and variation aims to provide a basis for a quantitative and objective expression and representation of prevailing transmissivity in hydrogeological maps and studies. Expressing transmissivity classes in a compact form enables a quick and objective comparison of different areas and/or hydrogeological environments in local and regional studies. Hydrogeological background and anomalies (positive and negative ones) can be determined if statistical treatment of samples is possible. Future water well results can then be antiticipatcd on the basis of a statistical approach. This methodology, of course, cannot replace detailed and com• plex mathematical models. It can, however, help to formu- late general regularities and to prepare a way for definition of prevailing natural conditions. It may be considered as the fiist step on the way towards hydrogcological data quanti• fication. Some interesting conclusions have been drawn from various studies that have applied the methods described in this paper: • Similarities or, on the contrary, significant differ- ences in prevailing transmissivity between distinct types of rocks or of whole hydrogeological basins [Carlsson and Carlstcdt, 1977; Krasny, 1975, 1976, 1986a; Krasny and L6pez, 1989; Michlicek, 1982; and many explanatory notes to hydrogeological map of Czechoslovakia I : 200,000, e.g., Hazdrova (ed.), 1983; Jetel (ed.}, 1986; Krasny (ed.), 1982]. • Differences in transmissivity depending on hydro- geological position of wells ( areas of recharge or discharge, Kdsny, 1974, 1984). • Vertical-depth permeability changes in different types of rocks (Krasny, 1975, 1986b). • On the basis of mean (prevailing) transmissivity values, a new method for ground-water runoff estimation in fissured "hard" rocks was proposed (Krasny and Knezek, 19n; Knezek and Krasny, 1990). A summary of some of the above-mentioned conclu- sions was published recently (K.rasny, 1990). After the first step of evaluating the spatial distribution of transmissivity and determining an arithmetic mean and a standard deviation, a redistribution of values among differ- ent statistical samples is sometimes useful; thus, more homogeneous samples can be produced, characterized by a narrower interval of the hydrogeological background x ± s and a less ambiguous classification oftransmissivity magni- tude with less transmissivity variation. Anomalies indicating zones with relatively greater or smaller transmissivity (per- meability) may give hints: revelation of buried alluvial fans may serve as an example (Krasny, 1983). In similar cases, hydrogeological conclusions may even influence geological conclusions. To maintain one of the principal advantages of the proposed classification, i.e., quick and easy data analysis, simplifying assumptions were accepted which usually do not affect results but which, however, should be taken into account. Some of them are as follows: l. Transmissivity (permeability) is considered to be a random function of a natural hydrogeological environment Therefore, the samples of data from wells are considered to be random samples from an infinite population. This is not the case for many available data samples, as many times water wells are purposefully located in more promising sites. As a consequence, the statistical distribution of a sample may be biased toward greater values; that is, the distribution is skewed more to the left when represented by a histogram. More realistic statistical models can be obtained when data are pretreated so that a selection of data from more perme- able zones is made by choosing only selected values as representative, or by calculation of a weighted arithmetic mean. 2. Unsuccessful water wells naturally should belong to a sample; however, usually neither specific capacity nor yield arc stated in archive reports. If the number of unsuc- cessful wells is known, an estimation of basic statistical characteristics is possible by graphical or numerical approx- imation of unknown value., by using the presumed statisti- cal, i.e. lognormal model distribution [Krasny (ed.), 1982]. In case of an unknown number offailures, no remedy exists and as a result, we have to consider the possibility that a sample where unsuccessful wells were probably omitted may overestimate the actual transmissivity. 3. Different depth of wells, naturally, may cause addi- tional variation of transmissivity values. That is why a pre- liminary selection should be made so as not to accept too great differences in depth of wells within the same statistical sample, particularly in hydrogeological environments where thickness of aquifer may influence transmissivity changes. Generally, however, relatively great differences in well depths (e.g., tens of meters when mean depth is about 100 meters) cannot significantly influence variation in transmis- sivity, as its horizontal changes caused by the variability within a geological environment (lithology, fracturing) are usually much more significant than those ones caused by different depth of wells. 4. Diameter of most drilled water wells ranges from 0.2 to 0.4 m. These differences do not usually influence statisti- cal characteristics significantly, as can be seen by examining equations for steady.,.tate radial flow toward a pumped well. Only large numbers of band-dug or other large-diameter wells such as arc bored in flu vial deposits or in the weathered zone of hard rocks may significantly influence the basic transmissivity characteristics of a sample. It is evident that these influences may cause some inaccuracies in transmissivity data generalization and regionalization. However, the author's experience during previous studies suggests that they do not affect the principal aim of the combined classification of transmissivity magni- tude and variation presented here: to quantify and standard- ize transmissivity expression, comparison, and representa- tion in bydrogeological maps and studies. Within samples of greater size, both inaccuracies and distinct influences are usually balanoed, and even relatively small differences in transmissivity magnitude and/or variation repeatedly found in some areas or environments might have logical hydrogco- logical grounds. Transmissivity classification has usually been used in regional hydrogeological studies, i.e. with most transmissiv- ity values measured at the scale of pumping tests from water wells and the space around them. Logarithmic-normal dis- tribution of statistical samples of permeability and transmis- sivity values prevail there. Transmissivity classification, however, could be used for studies of areas of different size (even under laboratory conditions) as well. In these cases, other types of statistical distribution (models) and also dif- ferent values could be expected because of the scale effect. Therefore, it should be explicitly stated, when using the classification, the extent of the studied area and conse- quently the method of transmissivity (permeability) assess- ment that was used. References Borevskiy, V. V., B. G. Samsonov, and L. S. Yazvin. 1979. Metod- ika opredcleniya paramctrov vodonosnykh gorizontov po dannym otkachek. Nedra, Moskva. 326 pp. Carlsson, L and A. Carlstedl 1977. Estimation of transmissibility and permeability in Swedish bedrock. Nordic Hydrology, Copenhlll!Cn, v. 8, pp. 103-116. Davis, S. N. and R.J.M. DcWicst. 1966. Hydrogcology. J. Wiley, New York-London-Sydney. 463 pp. Hazdrow, M. (ed.). 1983. Vysvctlivky k zllkladnl hydrogcologicke mape CSSR I :200 000, list 12 Praha (in Czech: Explana- tory notes to basic bydrogcological map of CSSR I :200,000 shcel 12 Praha). tlsuednl uslav gcologicky, Praha. 163 pp. !AH, !ASH, and UNESCO. 1983. International legend for 235 hydroseological maps. Revised version. 5 I pp. Jetcl, J. 1964. Pouzitl hodnot spccificke vydatnosti a novych odvozenych parametru vhydrogeologii (in Czech: Applica• tion of specific capacity values and new derived parameters in hydrogcology). Oeol. Prutlc., Praha. v. 6, no, 5, pp, 144-145, Jetel,J, (ed.), 1986. Vysvetlivky k zAldadnl hydrogeologickc mape CSSR l: 200 000, list 03 Liberec, 04 NAchod (cAst) [in Czech: Explanatory notes to basic hydrogcological map of CSSR l: 200,000, sheet 03 Liberec, 04 NAchod (part)]. tlstrednl ustav gcologicky, Praha. 157 pp. Jetel, J, and J. Knl.ffly. 1968, Approximative aquifer characteris- tics in rogional hydrogeological study, Vest. Ustr. Ust, gcol. Praha. v, 43, no, 5, pp, 459-461. Kiraly, L. 1975, Rappon surl-actucldes connaissanccsdans le domainc des caractUCS physiques des rochcs karstiqucs. In: Hydrogeology oflCarstic Terrains, !AH Paris. pp, 53~7. Knezek, M. and J. Krlasn)'. 1990. Natural groundwater resources mapping in mountainous areas of the Bohemian Massif. Mcm. 22nd Congress IAH. v, 2, Lausanne, pp. 846-854, Krasny, J. 1970, Zpracovanl a plosnc vyj!.dn:nl nektcrych hydro- gcologickych prvku v mapllch (in Czech: Elaboration and areal expression of some hydrogeological clements in maps). Shor. 5, hydrogcol. konfer., Zlln, pp, 57~. Krisn)', J. 1974. Les differences de la transmissivite, statistiquc- ment signilicatives, dans les zones d1nfihration et du drain- age, Mcm. AIH 10, I. Communic., Montpellier, pp, 204-211, Kr6snS,, J. 1975. Variation in tre.nsmissivity of L-rystalline rocks in southern Bohemia. Vest. tlstr, Ust. gcol. Praha. v, 50, no. 4, pp. 207-215. Krasny, J, 1976, Statisticka analS,,. hydrogcologickycb dat z podkrkonossk6 permokarbonskc panve (in Czech: Statisti- cal analysis of hydrogeological data from the Permo- Carboniferous Krkonosc-picdmont basin). Sbor. gcol. Vcd, Hydrogcol.-lnz, Geel. Praha. v, 13, pp, 113-1S2. Krasny, J. 1983. Hydrogcology and hydrochemistry of the Foot- hill Zone, MS 0.0. Geel. Survey and Miner, Investig,, Baghdad. Krasny, J, 1984. Vliv hydrogeologick6 po,ice homin na jejich propustnost (in Czech: Influence of hydrogcological posi- 236 tion ol rocks on their permeability). Geel. Pruzk,, Praha, v, 26, no, 12, pp, 342-34S. Kr6sny, J. 1986a. RegionAlnf tcndence hodnot transmisivity v jihomoravskcm fluviAlnlm kvan~ru (in Czech: Regional tendencies of transmissivity values in Southmoravian flu- vial Quaternary deposits), Zpr, gcol. Vyzk, 1984, tlstr. ust. gcol. Praha. pp. 121-123. Krasny, J, 1986b. Charakter propustnosti a transmisivita flyso- vych scdimentu jiznl casti Zd!nick6ho lesa (in Czech: Character of permeability and transmissivity of flysh deposits of the southcm pan of the Zdanicky les Mts.). Zpr, geol. Vyzk. 1984, tlstr. ust. gcol. Praha. pp, 119-121. Krjsny,J, 1986c. Klasifikaa: transmisivity ajcjl pou,.itf (in Czech: Transmissivity classification and its application). Gcol. Pruzk,, Praha. v, 28, DO, 6, pp, 177-179, KrAsnY. J. 1990. Regionali7.ation of transmissivity data: hard rocks of the Bohemian Massi!, Mem. 22nd Congress IAH, v. l, Lausanne, pp, 98-105. Krasny, J. (ed.), 1982. Vysvctlivky k zWadnl hydrogeologickc mape CSSR I: 200 000, list 13 Hradec Kralovc (in Czech: Explanatory notes to basic hydrogcologkal map ofCSSR l: 200,000, sheet 13 Hradcc KrAlovc). Ustrednl ustav gco- logicky, Praha, 159 pp. Krlt.snj, J. and M. Kncz.cl::. 19TI. Regional estimate of ground- water run-off from fissured rocks on using trtmsmissivity coefficient and geomorphologic characteristics. J. Hydro!. Sci., Warszawa. v, 4, no. 2, pp. 149-159. Kr6sny, J. and A, L6pcz, 1989, The general hydrogeological map of Nicaragua I: 250,000: Principles of compilation Wld content. Mem. Internal. Symp. on Hydrogcot maps as tools for econ. and social development, Hannover. pp. 413-422, Michllcek, E. 1982. Statistick! analyza transmisivity homin vy- chodni cAsti Cesk.omoravsk~ vrchoviny (in Czech: Statisti- ca1 analysis of transmissivity of rocks in the eastern pan of the Ccstomornvsk! vrchovina Mts.), Shor. gcol. Vcd, Hydrogcol.-Inz. Gcol. Praha. v. 16, pp, 91-120. Rats, M. V, 1967, Neodnorodnost gornykh pored i ikh fiziches• kith svoystv, Nauka, Moskva. Spiegel, M. R. 1972. Theory and Problems of Statistics in SI Units. McGraw-Hill Book Co, 3S9 pp.