• No results found

Effect of groundwater flow on forming arsenic contaminated groundwater in Sonargaon, Bangladesh

N/A
N/A
Protected

Academic year: 2021

Share "Effect of groundwater flow on forming arsenic contaminated groundwater in Sonargaon, Bangladesh"

Copied!
45
0
0

Loading.... (view fulltext now)

Full text

(1)

Effect of Groundwater Flow on Forming Arsenic Contaminated Groundwater in Sonargaon, 1

Bangladesh 2

3

Shinji Nakayaa, Haruyasu Natsumea, Harue Masudab, Muneki Mitamurab, Dipak Kumar Biswasc,

4

Ashraf A. Seddiqued

5 6

a

Department of Civil Engineering, Shinshu University, Nagano, Japan

7 b

Department of Geosciences, Osaka City University, Osaka, Japan

8 c

Asian Arsenic Network, Jessore, Bangladesh

9 d

Department of Petroleum and Mining Engineering, Jessore Science and Technology University, Bangladesh

10 11

Abstract 12

Three-dimensional groundwater flow in Sonargaon, Bangladesh is numerically simulated in order

13

to evaluate the flow paths of As-contaminated drinking groundwater in the Holocene aquifer of the

14

Ganges-Blamaptra-Meghna delta plain over a recent thirty-year period. The model indicates that

15

vertical infiltration of surface groundwater into the shallow Holocene aquifer occurs frequently in

16

the Ganges-Blamaptra-Meghna delta plain. It predicts that the water recharged from ground surface

17

moves approximately 10 m to 20 m vertically downward beneath the flood plain, with a gradually

18

increasing horizontal flow, toward the underlying Pleistocene middle mud layer (aquitard). The

19

model also predicts that groundwaters containing highest As concentrations (> 700 μg/L) are

20

formed on the vertical groundwater flow paths where surface water recharges the Holocene aquifer

21

and not on the horizontal flow paths. Combining with the groundwater chemistry, reducing

22

groundwater condition is not essential for the As-contaminated groundwater of the studied area in

23

the Ganges delta plain.

24 *Manuscript

(2)

25

Introduction 26

Large-scale natural arsenic groundwater contamination has been a serious problem in numerous

27

areas of the world, especially in Asian countries. In many cases it has a major impact on potable

28

water. Such contamination is typically associated with elevated pH in arid or semi-arid areas, or

29

strongly reducing conditions in geologically young sedimentary basins (e.g., Smedley and

30

Kinniburgh, 2002). In the most polluted area, approximately 35 million people in Bangladesh and 6

31

million in West Bengal, India are exposed to high levels of As (> 50 μg/L) in drinking water

32

(Geological Survey and Department of Public Health Engineering, 2001; Smedley and Kinniburgh,

33

2002; Nath et al., 2008). Elevated levels of groundwater As (> 50 μg/L) in this basin occur on a

34

large scale in strongly reducing alluvium-based aquifers at near-neutral pH, and in flat areas with

35

sluggish shallow groundwater flow. The mechanism and triggers of As dissolution from minerals to

36

groundwater has been addressed by extensive research, and the microbial reduction and dissolution

37

hypothesis, in which As enriched Fe-oxyhydroxides/oxides are decomposed and release the As into

38

the aquifer under reducing groundwater conditions, is widely accepted (e.g., Nickson et al., 2000;

39

McArthur et al., 2001; Bhattacharya et al., 2002; Smedley and Kinniburgh, 2002; van Geen et al.,

40

2003, 2004; Horneman et al., 2004; Zheng et al., 2005). In more recent studies, chemical

41

weathering of detrital basic minerals such as biotite in the aquifer, due to the enforced infiltration of

42

surface water, has been suggested as an essential mechanism for releasing As into the aquifer (e.g.,

(3)

Seddique et al., 2008; Itai et al., 2008).

44

In previous studies, it has been well documented that As-contaminated and As-free/less

45

contaminated areas are distributed as patches in close proximity to each other (e.g., Smedley and

46

Kinniburgh,2002; van Geen et al., 2003; Ravencraft et al., 2002; Mitamura et al., 2008; Itai et al.,

47

2008). Numerous researchers previously considered that the highly contaminated As groundwater

48

appeared in stagnant regions of the aquifer and that dissolved As was flushed out of deep aquifers

49

but not from the shallow aquifers, due to low groundwater mobility. Harvey et al. (2002; 2003),

50

however, documented that the circulation of local groundwater occurred on a scale of a few hundred

51

meters in diameter, and that such small-scale circulation activated in this thirty-year period

52

promoted As release into the groundwater. These results are considered valuable because it

53

revealed that the groundwater flow system was an essential factor causing As groundwater

54

contamination in the Ganges delta plain (e.g., Geological Survey and Department of Public Health

55

Engineering, 2001; Harvey et al., 2006; Klump et al., 2006; Mukherjee et al., 2007; Michael and

56

Voss, 2009; Neumann et al., 2010; Mukherjee et al., 2011). However, few studies focusing on the

57

relationship between the distribution of varying As contamination levels and groundwater systems

58

have subsequently been published.

59

By conducting a series of studies in Sonargaon, Bangladesh, we realized that the vertical

60

infiltration of surface water into the shallow aquifer promoted As contamination. In those studies,

61

Mitamura et al. (2008) reported that wells installed into fine sediments were highly contaminated by

(4)

As (highly As-contaminated wells were occasionally installed into fine micaceous sediments), and

63

that the geological structure of the aquifers was an important factor controlling the formation of

64

As-contaminated groundwater in Bangladesh. Seddique et al. (2008) reported that detrital biotite

65

was a primary source of As and that chemical weathering of this mineral was an essential

66

mechanism affecting the chemical composition of groundwater, including the As concentration

67

thereof (Itai et al., 2008). Heterogeneous distribution of major chemical components and the

68

isotopic ratios of oxygen and hydrogen in the groundwaters indicated vertical infiltration of surface

69

groundwater into the shallow sediments and short groundwater recharge paths (Itai et al., 2008). In

70

this study, three-dimensional transient groundwater flow is simulated with realistic assumptions of

71

hydraulic constants and boundary conditions of the geological structure, focusing on the

72

relationship between the flow paths and the concentrations of As in contaminated drinking

73

groundwater over a recent thirty-year period in the shallow groundwater system of the Holocene

74 aquifer. 75 76 Method 77

Transient three-dimensional groundwater flow was simulated using a numerical model to

78

document the flow paths and the residence time of As-contaminated groundwater in the Holocene

79

aquifer of the study area over thirty years. The model was constructed based on a data set including

80

topography; geological structure; hydraulic constants; tube well information (numbers, locations

(5)

and screen depths of wells drilled less than 40 m deep that are used mainly for drinking water); the

82

population living at each levee where there are settlements; transient water head records for three

83

observation wells installed at three different depths at the same site; temporal groundwater level

84

records at eight hand-auger drilling sites; irrigation; precipitation; evapotranspiration; and assumed

85

boundary conditions of aquifers surveyed in 2003 through 2006. The ground surface of all study

86

area can be recharge area and the river zone can be recharge or discharge area depending on

87

hydraulic gradient between river water head and groundwater head. Of the data used in this study,

88

the temporal groundwater level changes and geochemical characteristics such as major and minor

89

compositions, H, O, N and S isotopes, tritium units, etc., were already published in the previous

90

reports (Mitamura et al., 2008; Itai et al., 2008).

91

92

Site description and hydrogeological setting 93

The study area is 3.0 km × 3.0 km in size, located 20 km east of Dhaka, in Sonargaon,

94

Bangladesh. The area is composed of terraces in the west and a flood plain, including natural levees

95

along the old Brahmaputra River, which is an abandoned channel of the present Meghna River.

96

Except for the natural levees, the flood plain is inundated during the rainy season. During the dry

97

season, the floodplain dries and is used for cultivation, and the channel becomes narrower and

98

consists of a series of disconnected ponds (Fig. 1). The flat surface of the ground, which only

99

varies 4 m to 5 m in altitude, and the meandering river channel suggest a very low hydraulic

(6)

gradient and/or sluggish groundwater flow beneath the flood plain. Drinking water is drawn from

101

groundwater through tube wells installed into shallow and deep aquifers in the populated western

102

terrace and levees. Aquifers less than 90 m deep in the study area consist of upper and lower sand

103

formations separated by a mud layer; the upper Holocene sand formation is commonly 25 m to 35

104

m thick and overlies the Upper Pleistocene mud formation, which unconformably overlies the lower

105

Plio-Pleistocene sand formation (Mitamura et al., 2008) (Fig. 1). The upper sand formation thus

106

comprises the shallow unconfined aquifer, and the lower formation is the deep confined aquifer. The

107

upper sand formation hosts thin intercalated clay to silt layers in the upper section, and medium

108

sand in the lower section. The upper sand formation can, therefore, be divided into the uppermost

109

sand and the upper sand layers at approximately 0 m altitude. The mud formation, which functions

110

as an aquitard, also includes intercalated lenses of very fine sand to silt. The lower sand formation is

111

exposed as terraces in the western part of the study area. The upper sand formation thus abuts the

112

lower sand formation with an inferred fault along the boundary between the terraces and the alluvial

113

flood plain. The mud and lower sand layers are partly eroded to form valleys in this area.

114

115

Model description 116

Three-dimensional groundwater flow in unsaturated-saturated porous media is modeled by the

117

governing equation as follows:

118

t S C q K K x K K x ij i                    s 3 r r( ) ( ) ( ) ; i, j = 1, 2, 3 (1) 119

(7)

where  is hydraulic head, xi is a spatial coordinate, t is time, Kij is the saturated hydraulic

120

conductivity,  is the volumetric water content (= nSr), n is porosity, Sr is the saturation index 121

(ranging from 0 to 1), Kr(θ) is the unsaturated-saturated hydraulic conductivity ratio (also ranging 122

from 0 to 1), C() is the specific water content (=d/d), Ss is the specific storage, and  is a 123

function with a value of 1 when Sr = 1 and 0 when Sr < 1. 124

Equation (1) is solved for unknown  under appropriate boundary conditions. In the present study,

125

because unsaturated-saturated seepage flow including rain, evaporation and irrigation recharge in

126

unconfined aquifer is analyzed, the unsaturated-saturated three-dimensional seepage flow analysis

127

code AC-UNSAF3D (developed by Okayama University, Japan in 1980s and opened as a standard

128

code since 2000 in Japan: http://gw.civil.okayama-u.ac.jp/gel_home/download/index.html) was

129

applied with a three- dimensional finite element method (3D-FEM) scheme (Segol, 1977; Frind and

130

Verge, 1978; Huyakorn et al., 1986; Nishigaki et al., 1992) to document the three-dimensional

131

groundwater flow in this study area using a one month time step. This code is specified for

132

advective flow only.

133

134

Model ofstratigraphic structure and hydraulic properties

135

In the present FEM model, a region 3500 m in the x direction, 3000 m in the y direction and

136

114 m in the vertical (z) direction is divided into 106567 rectangular voxels of 100 m × 100 m ×Δz

137

[m] (Table 1). Because a rectangular grid system with small Δz/Δx and Δz/Δy ranging from

(8)

5x10-3 to 2x10-2 was adopted, FEM was used instead of finite difference method from the point of

139

analytical precision. The length in the x direction (3500 m) was extended 500 m eastward from the

140

study area to avoid boundary effects. Elements above the ground surface are excluded from the

141

model. To construct the model, the hydrogeological structure of the area is divided into three layers;

142

the upper sand formation, of which the uppermost S1 and S2 layers were defined from sediment

143

particle size, the lower sand formation S3, and the mud formation M, which functions as an aquitard

144

separating the upper and lower sand formations. The coordinates (x, y, z) of the boundary between

145

the aquifers and aquitard are defined based on lithostratigraphic data (Mitamura et al., 2008) of

146

sediment columns at 12 drilling sites by linear interpolation using the Surfer code (Golden Software,

147

Inc.) considering well depth of the 230 tube wells. For evaluating the volume of well water

148

withdrawn, well depths are assumed to be the same as the screen depths of the 230 tube wells we

149

studied. Figure 2 shows the contour map of the ground surface from SRTM 90 m digital elevation

150

data originally produced by NASA and the estimated geological structures. Depressions or buried

151

valleys, which partly eroded the mud formation, appeared as two WNW–ESE furrows and an

152

intermittent NNE-SSW furrow beneath the flood plain. One of the furrows along the WNW–ESE

153

direction, which includes drill site DRK-D1 (Figs. 1 and 2(e)) lacks the middle mud formation, and

154

the upper and lower sand formations connect directly.

155

Hydraulic properties determined by in situ pumping tests (Mitamura et al., 2008) were assigned

156

to each element of the model according to the boundary surfaces between S1, S2, S3, and M (Table

(9)

2). The hydraulic constants of aquitard M, the specific storage Ss, and the nonlinear hydraulic 158

properties of suction potential (– relationship) and hydraulic conductivity (Kr– relationship) for 159

each sedimentary formation are defined without hysteresis in the wetting and drying curves in the

160

unsaturated zone of the aquifers (Bear, 1972; Marsily, 1986; Huyakorn et al., 1986; Nishigaki et al.,

161

1992). If an element of the model is divided by the layer boundary surfaces, weighted average

162

values of the hydraulic constants are assigned to the element according to the volume occupied by

163

each aquifer or aquitard.

164

165

Simulation conditions 166

Under the hydrogeological setting described above, groundwater flow in the FEM model was

167

primarily determined for a thirty-year period simulating the conditions associated with temporal

168

changes in the level of the water table; precipitation, evapotranspiration, and groundwater pumping

169

rates for daily use and irrigation. The rate of groundwater pumping for daily use was estimated

170

based on data gathered in 2005 on the number of families, population, number of tube wells, and

171

well screen depth in each village (Table 3). The average monthly precipitation during 2005 in

172

Bangladesh (Fig. 3) was assigned to the nodal points on the ground surface. The floodwater table on

173

the flood plain along the old Brahmaputra River was assigned to nodal points on the ground surface,

174

set to vary with temporal changes of the groundwater potential observed in 2005 in the shallow

175

aquifer (upper sand formation) in wells DRK-W2(-18 m to -27 m) and DRK-W3 (-6 m to -12 m)

(10)

(Fig. 4). The monthly evapotranspiration (Fig. 5) was assigned to nodal points on the ground

177

surface estimated from the average monthly temperature between 1957 and 1987 at the Rangamati

178

meteorological observatory (22.63°N, 92.20°E, 63 m above sea level (ASL)), Bangladesh, using the

179

Thornthwaite equation (Thornthwaite, 1948). Fresh water withdrawal for domestic use in

180

Bangladesh has been estimated to be 13 L/(personday) in 1987 and 49 L/(personday) in 2000

181

(The World‟s Water; http://www.worldwater.org/data.html). 100 L/day/person was assumed as a

182

maximum value for recent withdrawals from the tube wells based on these figures. The Bangladesh

183

Rehabilitation Assistance Committee Office in Narayanganj, Bangladesh estimated well water

184

usage of 4 L/(personday) for this area (pers comm.). Thus, a constant pumping rate of either

185

4 L/(personday) or 100 L/(personday) was assigned to nodal points at the well screen positions at

186

5 m vertical intervals. The volume withdrawn from each well was calculated from the pumping rate,

187

local population, and numbers of tube wells, with flow percentages attributed to each well in each

188

village by depth (Table 3). Withdrawal of groundwater from the deep aquifer and recharge via its

189

distribution as much as the withdrawal onto the ground surface for irrigation during the dry season

190

were also taken into account in the numerical simulation using data for the 2003-2004 irrigation

191

season (January to April) in Munshiganj, Bangladesh (Harvey et al., 2006). The withdrawal from

192

the deep aquifer for seasonal irrigation (January to April) as much as the α percent of the

193

withdrawal (Qw) given by Harvey et al. (2006) (Fig. 3) and the seasonal recharge promoted by 194

irrigation as much as the α percent of the Qw were assigned to the nodal points at z = -70 m ASL 195

(11)

beneath the flood plain and at the ground surface, excluding levees, respectively, for the period from

196

January to April. In the present study, α was determined to be 75 % in the calibration by trial and

197

error as described in the chapter „Result‟. The volume of irrigation water taken from the deep

198

aquifer is much greater than that withdrawn for drinking water in this area; the water recharged due

199

to irrigation during the dry season is a maximum of 45 mm/month (The World‟s Water;

200

http://www.worldwater.org/data.html). The boundary conditions of inflow rate and potential head at

201

the ground surface nodes for the thirty modeled years were, therefore, controlled by monthly

202

changes in precipitation, evapotranspiration, irrigation groundwater recharge and the level of the

203

water table. As described before, horizontal groundwater flow direction will be macroscopically

204

from north to south and horizontal hydraulic gradient will be very low due to the flat surface and the

205

meandering river channel. Considering the temporally large changes of groundwater head at wells

206

and river water head in delta plain, the synchronically systematic potential head difference between

207

upstream and downstream must cause constant, horizontal groundwater flow. However, it is

208

difficult to obtain the hydrological data on the macroscopic, horizontal groundwater flow from

209

upstream to downstream using the only data within this study area. To consider the groundwater

210

flow from upstream to downstream, constant inflow and outflow rates were estimated to be the flow

211

rates equivalent to a hydraulic gradient of 1 m/10,000 m. The rates were estimated from

212

synchronically systematic potential head differences between 3.6 m and 7.7 m in this study area and

213

1.5 m and 5.7 m in Munshiganj (Harvey et al., 2006), and were assigned to the saturated nodal

(12)

points on the north and the south vertical boundary faces, respectively. The transient

215

three-dimensional groundwater flow analysis was carried out by a month as a time step.

216

217

Results 218

First, calibration or sensitivity analyses were carried out by checking the concordance between

219

calculated and observed results shown in Figures 6 and 7 to indicate the applicability of the

220

simulation model. Figures 6(a) and 6(b) show the numerically simulated groundwater head for the

221

1st to 15th years (a) and 16th to 30th years (b) in the 30-year simulation of wells DRK-W1,

222

DRK-W2, and DRK-W3 for groundwater usages of 4 L/(personday) and 100 L/(personday).

223

Figures 6(c) and 6(d) show the variation of simulated groundwater head in the 30th year for the 4

224

L/(personday) and 100 L/(personday) groundwater usage levels, respectively, which are the same

225

as the simulations of the 2nd to 29th years of the thirty-year simulation (Figs. 6 (a) and 6(b)) at

226

wells DRK-W1, DRK-W2, and DRK-W3 (Fig. 1). In Fig. 3, the irrigation rate is reduced to 75

227

percent of the initial value given by Harvey et al. (2006). The groundwater head variations obtained

228

are consistent with observed values to within 0.4 m. When omitting pumping for irrigation, the

229

computed groundwater heads in both the shallow (S1 and S2) and deep (S3) aquifers give the same

230

temporal variation, and the difference between the observed and calculated groundwater heads is 1

231

m to 1.5 m in the late dry to early rainy seasons from January to June. It is clear that the uptake of

232

groundwater for irrigation causes the groundwater head in the deep aquifer to be lower than in the

(13)

shallow aquifers in the dry season. The volume of groundwater withdrawn from tube wells for daily

234

use (4 L/(personday) or 100 L/(personday) does not affect the groundwater head throughout the

235

year, suggesting that pumping from the tube wells has a negligible effect on the groundwater flow

236

regimen in the study area. The simulated annual variation pattern of the groundwater head did not

237

change for 29 years, from the second year of the simulation, suggesting a recurrent hydrological

238

cycle in the area. Stability of the simulation results under the present assumptions and boundary

239

conditions are also confirmed from the periodicity for those 29 years. Figure 7 shows a map of

240

unconfined groundwater head computed for October, early in the dry season, with the observed

241

levels at eight hand-auger drilling sites in early October, 2006. Despite uncertainties in the

242

hydrologic data, the spatial variation of simulated groundwater heads are concordant with the

243

observed levels, indicating that the model used here appropriately simulates the groundwater flow

244

regimen for this study area.

245

Figure 8 shows simulated monthly horizontal groundwater flow velocities at Z = -10 m from

246

January to December. The horizontal groundwater flow differs markedly between the dry and rainy

247

seasons. In the dry season, from January to May, the dominant groundwater flow in the shallow

248

aquifer (S2) is around and toward a WNW–ESE furrow in the southeastern part of the study area,

249

where the middle mud formation is thin or non-existent (Fig. 2(e)). The groundwater flows from the

250

shallow aquifer into the deep aquifer by contact between the aquifer sediments due to the lack of the

251

middle mud formation (Figs. 8(a) to (e) and 2(e)), indicating that irrigation pumping affects the

(14)

groundwater flow in the dry season. Between July and September in the rainy season, horizontal

253

flow is limited beneath the flood plain, and its dominant pattern appears to be radial flow from the

254

levees to surrounding areas (Figs. 8(g), (h) and (i)). Thus the low horizontal hydraulic gradient of

255

the region increases the extent of stagnant conditions in the aquifer during the period of annual

256

flooding. The radial groundwater flow, typically in actively recharge zones, is produced by the

257

positive groundwater potential of the large rainy season precipitation to raise the water table of the

258

shallow aquifer. The village of Harihardi (Fig. 1) is such an active recharge zone. From October to

259

November, early in the dry season, the stagnant zones become small, and the velocity of the radial

260

flow increases (Figs. 8(j) and (k)). In June and December, the groundwater flow changes between

261

the dry and rainy season patterns (Figs. 8(f) and (l)). The simulation also shows slight groundwater

262

flows from the terrace into the deep aquifer (S3) beneath the flood plain in both dry and rainy

263

seasons.

264

To evaluate the groundwater residence, the groundwater flow path toward the tube well was

265

traced by forward and backward particle tracking of the particle source placed at an arbitrary point,

266

assuming advection of hypothetical, non-reactive particle (e.g., Mukherjee et al., 2011) using

267

spatiotemporal distribution of the flow velocities computed by the three-dimensional groundwater

268

flow analysis. Figures 9(a) to 9(f) show the modeled groundwater flow paths toward the location of

269

tube wells Nos. 1 through 6 at Z= 5, 0, -5, -10, -15, -20, -25, -30 m. Water recharged from the

270

ground surface infiltrates between 5 m and 20 m vertically downward, with its horizontal flow

(15)

component increasing with depth, toward the drilled hole. The groundwater flow path toward the

272

well becomes longer and the horizontal component becomes larger with increasing depth. Figure

273

9(g) shows the relationship between z and residence time (model age) (TR) calculated for the wells 274

Nos. 1 through 6. In this study, the residence time (model age) means the time required for particle

275

to reach the target point from a recharge point. Although the simulation is limited to a thirty year

276

run, the obtained model age can be longer than thirty years because the model age can be calculated

277

from the sum of TR on arbitrary segments along the groundwater flow path. The model age of 278

groundwater in the Holocene aquifer increases with depth and is proportional to depth down to

279

approximately -20 m. The correlation between depth, as it increases to about 20 m below the ground

280

surface, and the model age of the groundwater is concordant with the correlation between the depth

281

and residence time of groundwater estimated from 3H/3He isotopes in the Holocene aquifers of the

282

Munshiganji District (Klump et al., 2006) and in Araihazar Upazila (Stute et al., 2007).

283

Figures 10 and 11 demonstrate the simulated groundwater flow paths toward the midpoint of the

284

5-meter long screen, along with As concentrations for the 126 surveyed tube wells installed in the

285

shallow (Holocene) aquifer (above about -20 m ASL). The present model predicts that vertical

286

infiltration of surface water into the shallow aquifer sediments occurs often; the water recharged

287

from the ground surface (about 4 m to10 m ASL) of the flood plain moves approximately 5 m to 20

288

m downward and the groundwater flow gradually changes toward the horizontal direction as it

289

approaches the underlying middle mud layer (aquitard M). The drinking water pumped up is

(16)

derived from the groundwater flowing downward from the ground surface toward the tube well. As

291

noted above, recharge and subsequent radial flow appear in and around levees. A stagnant zone

292

appears in the rainy to early dry season, July to October (figs. 7-11), while recharge is promoted in

293

the stagnant zone in the dry season. The model also predicts that the shallow groundwater in the

294

Holocene aquifer will not recharge in the terrace (figs. 10). The areas with the highest As

295

concentrations (700 μg/L < As < 1200 μg/L) and high As concentrations (500 μg/L < As < 700

296

μg/L) correspond to the zones where recharge occurs in the rainy to early dry seasons (Figs. 7, 8

297

and 10). Thus, local groundwater flow and the As concentration in the flow path must be important

298

for the mechanism that produces As-contaminated groundwater. Figure 12 shows the spatial

299

distribution of As and the concentration of dissolved oxygen (DO) , which was measured after

300

attainment of the stabilities of DO, ORP, EC and pH with enough pumping, of 45 of the total 51

301

points measured from the end of September to early November in 2009 in eastern levees where an

302

active recharge zone appears. DO is detected in 20 of the 51 well samples. The most active recharge

303

zone (Figs. 8, 10 and 11) in this study area appears in the village of Harihardi village (Fig. 1), where

304

DO levels of 1.3 mg/L to 3.3 mg/L were detected in 18 of the 28 wells. Two well samples

305

containing fairly amounts of DO (1.73 and 1.76 mg/L) relative to DO (0 mg/L) in reducing

306

condition were among the seven that had As > 500 μg/L.

307

Figure 13 shows computational results presenting the relationship between the average length of

308

the groundwater flow path (Lp) and the average residence time (model age) (TR) from the ground 309

(17)

surface to tube wells, with the As concentration of each well‟s water, excluding three wells that had

310

residence times >30 years (920 m, 38.5 years, 351 As μg/L), (367 m, 43 yeas, 146 As μg/L) and

311

(532 m, 31 years, 75 As μg/L). The estimated Lp ranges from 38 m to 920 m, and TR ranges from 4 312

years to 43 years. The TR obtained here is concordant with the estimation that the Holocene aquifer 313

(Mitamura et al., 2008) hosts groundwater recharged after 1953 based on 3H measurements of this

314

study area (Itai et al., 2008). Groundwaters in the As hotspots (> 700 μg/L) are characterized by

315

short flow paths and residence times; the hotspots with the highest As concentrations (> 700 μg/L)

316

have flow paths shorter than 220 m and residence times (model ages) <10 years.

317

318

Discussion 319

Here, the As contamination process is discussed in relation to simulated flow paths for a recent

320

thirty-year period in the Holocene aquifer. As noted above, shallow groundwater in the Holocene

321

aquifer has a downward flow path from the ground surface toward each tube well. The sources of

322

that groundwater are rivers, local rainwater, and a small contribution of deeper groundwater

323

distributed on the cultivation field for the irrigation. As groundwater hotspots in this area are closely

324

related to the groundwater flow path after the recharge: our simulation shows that the among the

325

wells we studied As hotspots appear in wells having relatively short Lp and TR as shown in Fig. 13. 326

Figure 14 shows two conceptual models, A (CM-A), in which the As is released mainly during

327

vertical groundwater infiltration, and B (CM-B) in which the As is released over the entire flow

(18)

path where As is concentrated in the aquifer sediments. Through the groundwater flow, advection,

329

dispersion and reactive transport of As often occur. In case CM-A (Fig. 14(a)), the As released in the

330

vertical flow path would be diluted by the ambient horizontally flowing groundwater. Even if the

331

groundwater passes through As-rich sediment, little As is released. Therefore, As hot spots appear

332

close to the recharge zones in places where the groundwater has short flow paths and residence

333

times. This model can explain the relationship between the TR and the Lp of the groundwaters in the 334

Holocene aquifer shown in Fig. 13. The linear correlation between groundwater age determined

335

using 3H/3He, at depths <20 m and the dissolved As concentration was observed in Araihazar

336

Upazila (Stute et al., 2007). These authors also found that the age of groundwaters containing the

337

highest As concentrations > 1100 µg/L (hotspots) are considerably younger than the others, of

338

which ages are plotted on the linear regression line. Thus, the considerable As release in Araihazar

339

Upazila would occur in the vertical flow path through As-rich sediments, which are distributed as

340

patches, similar to our model CM-A. In case CM-B (Fig. 14(b)), As is not only released in the

341

vertical flow path, but also in the following horizontal path. According to this model, high As

342

concentrations in the groundwater appear to be independent of the length of the flow path and

343

residence time. Therefore, case CM-B cannot explain the As hotspots of this study area. We

344

conclude that the release of As from the Holocene aquifer in our study area is likely to occur

345

primarily in vertically infiltrating water. Arsenic contaminated groundwater in Asian countries is

346

typically found under two different groundwater conditions (e.g., Smedley and Kinningurgh, 2002);

(19)

i.e., reducing conditions in humid climates, such as Bangladesh, Vietnam, etc., and oxic conditions

348

in arid to semi-arid climates, such as Pakistan and Mongolia. The formation mechanism in reducing

349

groundwater is believed to be microbial reduction and decomposition of Fe-oxyhydroxides,

350

releasing As, which is adsorbed onto the Fe-oxyhydroxides (e.g., Nickson et al., 2000; Horneman et

351

al., 2004; van Geen et al., 2004; Zheng et al., 2004; Mukherjee et al., 2008). The As release in the

352

oxic groundwater is explained by desorption of As from Fe-oxyhydroxides without decomposition

353

(e.g., Amini et al., 2008). According to these hypotheses, the As in the groundwater of the oxic

354

recharge zone in the study area is likely released via desorption from Fe-oxyhydroxides without

355

decomposition. Some researchers propose silicates as sources of the As; e.g., biotite has been

356

considered to be a primary source (Seddique et al., 2008), although chlorite was demonstrated to

357

enrich As in Masuda et al. (2010). Such basic minerals are easily decomposed via chemical

358

weathering, which is promoted by water flow irrespective of the redox condition, although an oxic

359

condition is preferable to an anoxic condition. In our study area, redox condition of the Holocene

360

groundwaters allows to precipitate FeOOH (goethite) and Fe(OH)3 (ferrihydrite) (Itai et al., 2008). 361

They also documented that the chemical weathering of detrital minerals such as plagioclase and

362

basic minerals are important factors to form chemical composition of the groundwaters. If chemical

363

weathering of As-bearing mineral(s) is the case, higher As concentrations can be expected in the

364

rather oxic recharge zone where the groundwater flow rate is higher than that in the stagnant anoxic

365

zone. This study cannot specify the As contamination mechanism at present, but it is clear that a

(20)

reducing groundwater condition is not essential to form As contaminated groundwater of the study

367

area, and presumably even for the other As contaminated groundwater in the Ganges delta plain.

368

369

Acknowledgements 370

The authors would like to thank anonymous reviewers, Associate Editor and Dr Laurent Charlet

371

for constructive reviews and helpful comments. We would like to thank Dr Kazi Matin Ahmed for

372

assistance and support on the field survey. We would also like to thank Ataur Rahman (BRAC) for

373

additional information of well and population for each village in the study area. This research was

374

partly supported by a Grant-in-aid for Science Research, number 15403017, from the Japanese

375

Ministry of Education, Science, Sports and Culture.

376

377

References 378

Amini M., Abbaspour, K. C., Berg, M., Winkel, L., Hug, S. J., Hoehn, E., Yang, H., Johnson, C. A.,

379

2008. Statistical Modeling of Global Geogenic Arsenic Contamination in Groundwater. Environ.

380

Sci. Technol., 42 (10), 3669–3675, DOI: 10.1021/es702859e.

381

Bear, J., 1972. Dynamics of fluids in porous media, American Elsevier Pub. Co., New York.

382

Bhattacharya, P., Jacks, G., Ahmed, K. M., Routh, J., Khan, A. A., 2002. Arsenic in groundwater of

383

the Bengal Delta Plain Aquifers in Bangladesh. Bull. Environ. Contam. Toxicol., 69, 538-545,

384

DOI: 10.1007/s00128-002-0095-5.

(21)

British Geological Survey and Department of Public Health Engineering, Arsenic Contamination of

386

Groundwater in Bangladesh, vol. 2, Final Report, BGS Tech. Rep. WC/00/19, edited by D. G.

387

Kinniburgh and P. L. Smedley, Br. Geol. Surv., Keyworth, UK, 2001.

388

Frind, E. O., Verge, M. J., 1978. Three-dimensional modeling of groundwater flow system. Water

389

Resour. Res., 14(5), 844-856.

390

Harvey, C. F., Ashfaque, K. N., Yu, W., Badruzzaman, A. B. M., Ali, M. A., Oates, P. M., Michael,

391

H. A., Neumann, R. B., Beckie, R., Islam, S., Ahmed, M. F., 2006. Groundwater dynamics and

392

arseniccontamination in Bangladesh. Chemical Geology, 228, 112-136.

393

Harvey, C. F., Swartz, C. H., Badruzzaman, A. B. M., Keon-Blute, N., Yu, W., Ali, M. A., et al.

394

2002. Arsenic mobility and groundwater extraction in Bangladesh. Science 298(5598),

395

1602–1606.

396

Harvey, C. F., Swartz, C. H., Badruzzaman, A. B. M., Keon-Blute, N., Yu, W., Ali, M. A., et al.

397

2003. Response to comments on “Arsenic mobility and groundwater extraction in Bangladesh”.

398

Science 300(5619), 584.

399

Horneman, A., van Geen, A., Kent, D. V., Mathe, P. E., Dhar, R. K., O'Cinekk, S., Hoque, M. A.,

400

Aziz, Z., Shamsudduha, M., Seddique, A. A. Ahmed, K. M., 2004. Decoupling of As and Fe

401

release to Bangladesh groundwater under reducing conditions. Part I: Evidence from sediment

402

profiles. Geochim. Cosmocim. Acta, 68(17), 3459-3473.

403

Huyakorn, P. S., Springer, E. P., Guvanasen, V., Wadsworth, T. D., 1986. A three dimensional

(22)

finite-element model for simulating water flow in variably saturated porous media. Water

405

Resour. Res., 22(13), 1790-1808.

406

Itai, T., Masuda, H., Seddique, A.A., Mitamura, M., Maruoka, T., Li, X., Kusakabe, M., Dipak, B.K.,

407

Farooqui, A., Yamanaka, T., Nakaya, S., Matsuda, J., Ahmed, K.M., 2008. Hydrological and

408

geochemical constraints on the mechanism of formation of arsenic contaminated groundwater in

409

Sonargaon, Bangladesh. Appl. Geochem. 23, 3155-3176.

410

Klump, S., Kipfer, R., Cirpka, O. A., Harvey, C. F., Brennwald, M. S., Ashfaque, K. N.,

411

Baduruzzaman, A. M., Hug, S. H., Imboden, D. M., 2006. Groundwater dynamics and arsenic

412

mobilization in Bangladesh assessed using noble gases and tritium. Environ. Sci. Technol. 40,

413

243-250.

414

Masuda, H., Shinoda, K., Noguchi, N., Okudaira, T., Takahashi, Y., Mitamura, M., Seddique, A.A.,

415

2010. Chlorite as a primary source of arsenic in groundwater aquifer sediments in Bengal delta.

416

Geochim. Cosmochim. Acta 74 (12S1), A676.

417

McArthur, J. M., Ravencroft, P., Safiullah, S., Thirlwall, M. F., 2001. Arsenic in groundwater:

418

Testing pollution mechanisms for sedimentary aquifers in Bangladesh. Water Resour. Res.,

419

37(1), 109-117.

420

Michael H. M., Voss C. I., 2009. Controls on groundwater flow in the Bengal Basin of India and

421

Bangladesh: regional modeling analysis. Hydrogeology Journal, 17, 1561–1577, doi

422

10.1007/s10040-008-0429-4.

(23)

Mitamura, M., Masuda, H., Itai, T., Minowa, T., Maruola, T., Ahmed K. M., Seddique, A. A. Dipak,

424

B. K., Nakaya, S., Li, X. D., Uesugi, K., Kusakabe, M., 2008. Geological structure of an

425

arsenic-contaminated aquifer at Sonargaon, Bangladesh. J. Geology, 116, 288-302.

426

Mukherjee, A., Fryar, A. E., Howell, P. D., 2007. Regional hydrostratigraphy and groundwater flow

427

modeling in the arsenic-affected areas of the western Bengal basin, West Bengal, India,

428

Hydrogeology Journal, 15, 1397–1418, doi 10.1007/s10040-007-0208-7.

429

Mukherjee, A., Fryar, A. E., Scanlon, B. R.,Bhattacharya, P., Bhattacharya, A., 2011. Elevated

430

arsenic in deeper groundwater of the western Bengal basin, India: Extent and controls from

431

regional to local scale, Appl. Geochem., 26(4), 600-613.

432

Mukherjee, A., von Brömssen, M., Scanlon, B. R., Bhattacharya, P., Fryar, A. E., Hasan, M. A.,

433

Ahmed, K. M., Chatterjee, D., Jacks, G., Sracek, O., 2008. Hydrogeochemical comparison and

434

effects of overlapping redox zones on groundwater arsenic near the Western (Bhagirathi

435

sub-basin, India) and Eastern (Meghna sub-basin, Bangladesh) margins of the Bengal Basin, J.

436

Contaminant Hydrology, 99, 31-48, doi:10.1016/j.jconhyd. 2007.10.005.

437

Nath, B., Stüben, D., Basu Mallik, S., Chatterjee, D., Charlet, L., 2008. Mobility of arsenic in West

438

Bengal aquifers conducting low and high groundwater arsenic, part I: comparative

439

hydrochemical and hydrogeological characteristics. Appl. Geochem. 23, 977–995.

440

Nickson, R., McArthur, J. M., Ravenscroft, P, Burgess, W. G., Ahmed K. M., 2000. Mechanism of

441

arsenic release to groundwater, Bangladesh and West Bengal. Appl. Geochem. 15, 403-413.

(24)

Nishigaki, M., Nakaya, S., Kono, I., 1992. A three dimensional finite element method for seepage

443

through a saturated/unsaturated porous media by using SSOR-PCG method (Japanese with

444

English abstract). JSCE, -19(448), 101-110.

445

Neumann, R.B., Ashfaque, K.N., Badruzzaman, A.B.M., Ali, M.A., Shoemaker, J.K., Harvey, C.F.,

446

2010. Anthropogenic influences on groundwater arsenic concentrations in Bangladesh. Nature

447

Geosci. 3, 46–52.

448

Sakellariou-Makrantonaki, M., Tzimopoulos, C., Gouliaras, D., 1987.Analysis of a closed-form

449

analytical model to predict the hydraulic conductivity function. J. Hydrology, 92, 28-300.

450

Seddique A. A., Masuda, H., Mitamura, M., Shinoda, K., Yamanaka, T., Itai, T., Maruoka, T.,

451

Uesugi, K., Ahmed, K. M., Biswas, D. K., 2008. Arsenic release from biotite into a Holocene

452

groundwater aquifer in Bangladesh. Appl. Geochem. 23, 2236-2248.

453

Segol, G., 1977. A three-dimensional Galerkin-finite element model for the analysis of contaminant

454

transport in saturated-unsaturated porous media. in Finite Elements in Water Resources, edited

455

by W. G. Gray, G. F. Pinder, and C. A. Brebbia, 2123-2144, Pentech, London.

456

Smedley, P. L., Kinninburgh, D. G., 2002. A review of the source, behavior and distribution of

457

arsenic in natural waters. Appl. Geochem. 17, 517-568.

458

Stute, M., Zheng, Y., Schlosser, P., Horneman, A., Dhar, R. K., Datta, S., Hoque,M. A., Seddique, A.

459

A., Shamsudduha, M., Ahmed, K. M., van Geen, A, 2007. Hydrological control of As

460

concentrations in Bangladesh groundwater. Water Resour. Res., 43, W09417,

(25)

doi:10.1029/2005WR004499.

462

Thornthwaite,C.W., 1948. An approach toward a rational classification of climate. Geographical

463

Review, 38, 55-94.

464

van Geen, A., Rose, J., Thoral, S., Garnier, J. M., Zheng, Y., Bottero, J. Y., 2004. Decoupling of As

465

and Fe release to Bangladesh groundwater under reducing conditions. Part II: Evidence from

466

sediment incubations. Geochim. Cosmochim. Acta 68(17), 3475-3486,

467

doi:10.1016/j.gca.2004.02.014.

468

van Geen, A., Zheng, Y., Versteeg, R., Stute, M., Horneman, A., Dhar, R., Steckler, M., Gelman, A.,

469

Small, C., Ahsan, H., Graziano, J. H., Hussain I., Ahmed, K. M., 2003. Spatial variability of

470

arsenic in 6000 tube wells in a 25 km2 area of Bangladesh. Water Resour. Res. 3(5), 1140,

471

doi:10.1029/2002WR001617.

472

van Genuchten, M. Th., 1980. A closed-form equation for predicting the hydraulic conductivity of

473

unsaturated soils. Soil Sci. Am. J., 44(5), 892-898.

474

Zheng, Y., van Geen, A., Stute, M., Dhar, R., Mo, Z., Cheng, Z., Horneman, A., Gavrieli, I.,

475

Simpson, H.J., Versteeg, R., Steckler, M., Grazioli-Venier, A., Goodbred, S., Shahnewaz, M.,

476

Shamsudduha, M., Hoque, M.A., Ahmed, K.M., 2005. Geochemical and hydrogeological

477

contrasts between shallow and deep aquifers in two villages of Araihazar, Bangladesh:

478

implications for deeper aquifers as drinking water sources. Geochim. Cosmochim. Acta 69(22),

479

5203–5218.

(26)

Zheng, Y., Stute, M., van Geen, A., Gavrieli, I., Dhar, R., Simpson, H. J., Schlosser, P., Ahmed, K.

481

M., 2004. Redox control of arsenic mobilization in Bangladesh groundwater, Applied

482

Geochemistry, 19, 201–214, doi:10.1016/j.apgeochem.2003.09.007.

483

(27)

Figure captions 485

Figure 1. (a) Sampling wells and drill sites in study area. (b) Sediment columns at twelve drilling

486

sites. (c) Inferred geologic cross section (E-W). (d) Inferred geologic cross section (N-S).

487

Sedimentary layers are divided into upper and lower sand formations, and successive middle mud

488

formation separating between the upper and the lower sand formations. This figure is modified from

489

Mitamura et al. (2008).

490

491

Figure 2. Contour maps of the ground surface from SRTM 90m digital elevation data originally

492

produced by NASA (a), base of uppermost sand formation (aquifer 1) (b), base of upper sand

493

formation (aquifer 2) (c), base of middle mud formation (aquitard) (d) and thickness of the middle

494

mud formation (e) estimated from lithostratigraphic data (Figure 1) by linear interpolation.

495

496

Figure 3. Average monthly precipitation during 2005 and irrigation in Bangladesh. The averaged

497

precipitation was inferred from the data cited on the web site of Bangladesh Meteorological

498

Department (http://www.bangladeshonline.com/bmd/). The irrigation was obtained from data for

499

the 2003-2004 irrigation season (January to April) in Munshiganj, Bangladesh (Harvey et al., 2006).

500

501

Figure 4. Temporal change of the groundwater potential observed at well DRK-W1 in deep

502

aquifer (lower sand formation), and at wells DRK-W2 and W3 in shallow aquifer (upper sand

(28)

formation) in 2005. The screens of observation wells at Darikandi site(Figure 1) were installed in

504

45 to 54 m depth (W1), 18 to 27 m depth (W2), and 6 to 12 m depth (W3).

505

506

Figure 5. Monthly evapo-transpiration estimated from the average monthly temperature between

507

1957 and 1987 at the Rangamati meteorological observatory (22.63°N, 92.20°E, 63 m above sea

508

level (ASL)), Bangladesh, using the Thornthwaite equation (Thornthwaite, 1948).

509

510

Figure 6. Numerically simulated groundwater head for the 1st to 15th year (a) and 16th to 30th

511

years (b) in 30-years simulation of wells DRK-W1, W2, and W3 for groundwater usage of both 4

512

and 100 L/(personday). (c) and (d) show the variation of simulated groundwater head in the 30th

513

year for the 4 L/(personday) and 100 L/(personday) groundwater usage levels, respectively, with

514

observed values at wells DRK-W1, DRK-W2, and DRK-W3 (Fig. 1). H-Lower and H-upper

515

indicate the observed groundwater heads in lower and upper sand formation, respectively, in 2005.

516

Hcal indicates the simulated groundwater heads.

517

518

Figure 7. Contour map of unconfined groundwater head computed for October, early in the dry

519

season, with the observed levels at eight hand-auger drilling sites in early October, 2006.

520

521

Figure 8. Horizontal groundwater flow velocities simulated at Z=-10m in January (a), February (b),

(29)

March (c), April (d), May (e), June (f), July (g), August (h), September (i), October (j), November

523

(k) and December (l).

524

525

Figures 9. Modeled groundwater flow paths toward the location of tube wells Nos. 1 through 6 at

526

Z= 5, 0, -5, -10, -15, -20, -25, -30 m i. e. tube well of Nos. 1 (a), 2 (b), 3 (c), 4(d), 5 (e) and 6 (f). (g)

527

Relationship between z and residence time (model age) calculated for the wells Nos. 1 through 6.

528

529

Figure 10. Simulated groundwater flow paths toward the midpoint of the 5-meter long screen, along

530

with As concentrations for the 126 surveyed tube wells installed in the shallow (Holocene) aquifer

531

(above about -20m ASL). (a) As <300 mg/L, (b) As >300 mg/L.

532

533

Figure 11. Vertical profile of simulated groundwater flow paths toward the midpoint of the 5-meter

534

long screen, along with As concentrations for the 126 surveyed tube wells installed in the shallow

535

(Holocene) aquifer (above about -20m ASL). (a) x-z plane, (b) y-z plane.

536

537

Figure 12. Spatial distribution of As and the concentration of dissolved oxygen (DO) of 45 of the

538

total 51 points measured from the end of September to early November in 2009 in eastern levees

539

where a recharge zone appears.

540

(30)

Figure 13. Relationship between the average length of the groundwater flow path (Lp) and the

542

average residence time (model age) (TR) from the ground surface to tube wells, with the As 543

concentration of each well‟s water, excluding three wells that had residence times >30 years (920 m,

544

38.5 years, 351 As μg/L), (367 m, 43 yeas, 146 As μg/L) and (532 m, 31 years, 75 As μg/L). V

545

indicates velocity by Lp/TR. 546

547

Figure 14. (a) Conceptual model A (CM-A) and (b) conceptual model B (CM-B) for As release in

548

the groundwater flow system obtained in this study.

549

550

(31)

Tables 552

553

Table 1. Finite Element Size Using in 3D-FEM Analysis.

554 555 Depth (m) Δx (m) Δy (m) Δz (m) 556 557 14 – 0 100 100 0.5 558 0 – -70 100 100 1 559 -70 – -100 100 100 2 560 561 562 563

Table 2. Hydraulic Properties of Aquifers and Aquitard Using in 3D-FEM Analysis.

564

565

Soil layer K (m/sec) Ss (1/m) ne 566 567 Uppermost sand 7.4x10-5 10-7 0.3 568 Upper sand 2.6x10-4 10-7 0.3 569 Lower sand 1.6x10-4 10-7 0.3 570 Middle mud 1.0x10-8 10-4 0.6 571

Unsaturated property for sand:

572 Se = (Sr - Srr) / (1 - Srr), φ < 0 573 Se = 1, φ ≥ 0 574 Kr = Se 575 +Se =[ (1+(aφ)b ](1-b)/b, ++a=0.015(cm-1), ++b=5, Srr = 0 576

Where φ ; Pressure head (m), Sr ; Water saturation, 577

Srr ; Residual water saturation and ne ; effective porosity. 578 579 +: van Genuchten (1980), 580 ++: Sakellariou-Makrantonaki et al. (1987) 581 582 583 584 585 586 587 588

(32)

589

Table 3. Survey Data for Each Village of the Study Area in 2005.

590

Village name Family Population Number of Persentage of tube well (%) / depth

591 Tube well - 15.24m 15.24 - 30.48 - 45.72 - 60.96 - 91.44m – 592 30.48m 45.72m 60.96m 91.44m 593 594 Muchar Char (MCC) 303 2086 293 0 82 15 0 1 2 595 Harihardi (HHD) 74 379 73 0 85 9 6 0 0 596 Darikandi (DKD) 69 364 40 0 88 0 13 0 0 597 Dolardi (DLD) 169 1062 109 0 54 41 4 0 0 598 Kumarchar (KMC) 113 580 66 0 61 31 6 3 0 599 Gankulkandi (GLK) 65 317 38 0 88 13 0 0 0 600 Mamurdi (MMD) 116 631 68 0 97 0 0 3 0 601 Gulnagar (GLG) 102 536 74 0 43 43 0 0 0 602 Atbari (ABA) 38 226 19 0 100 0 0 0 0 603 Bara Kaitargaon (BKB) 193 1216 95 0 73 0 12 15 0 604 Ledamdi (LDD) 45 251 26 0 88 0 13 0 0 605 Temdi (TMD) 70 346 41 0 86 0 0 14 0 606 607

(33)

GLG-D1DLD-D 1 KMC-D1MCC-D1MCC-D2MCC-D3MCC-D4DRK-W1DRK-D 1 HHD-D 1 HHD-D 2 MMD-D1 GLG-D1 DLD-D 1 MCC-D4MCC-D3 HHD-D 1 MMD-D1 0 500 m 0 500 m W E S N MCC-D1 MCC-D2 MCC-D3 DRK-D 1 DRK-W1 0 -20 -40 -60 0 -20 -40 -60 Altitude (m) Altitude (m) 0 10 20 30 40 50 60 70 80 Depth (m) clay - silt silt - v.f. sand v.f. - f. sand f. sand f. - m. sand m. - c. sand c. sand Sampling well : Depth

Lower sand formation

Upper sand formation

Middle mud formation (a) (b) (c) (d) DRK-D1 Dhaka Study area Brahmaputra R. Ganges R. 88oE 90oE 92oE 26oN 24oN 22oN India India Bay of Bengal Old Brahmaputra R. Meghna R.

Fig. 1, Nakaya et al.

(34)

0 1000 2000 3000 0 1000 2000 3000 Y (m ) X (m) 0 1000 2000 3000 0 1000 2000 3000 Y (m ) 0 1000 2000 3000 0 1000 2000 3000 Y (m ) 0 1000 2000 3000 0 1000 2000 3000 X (m) 3.82 15 (m) 0 15 (m) -44 15 (m) -54 15 (m) Thickness 1 27 (m)

(a)

(b)

(c)

(d)

(e)

Fig. 2, Nakaya et al.

(35)

0

100

200

300

400

500

600

1

2

3

4

5

6

7

8

9

10

11

12

P

re

ci

p

it

a

ti

o

n

,

Ir

ri

g

a

ti

o

n

(m

m

/m

o

n

th

)

Month

Precipitation

Irrigation

Figure 3.Average monthly precipitation during 2005 and irrigation

in Bangladesh. The averaged precipitation was inferred from the

data cited on the web site of Bangladesh Meteorological

Department (http://www.bangladeshonline.com/bmd/). The

irrigation was obtained from data for the 2003-2004 irrigation

Fig. 3, Nakaya et al.

(36)

Jan.

Mar

.

May

.

Jul.

Sep.

Nov

.

Feb.

Apr

.

Jun.

Aug.

Oct.

Dec.

Time

0

1

2

3

4

5

Observed (2005)

W1 (45-54 m)

W2 (18-27 m)

W3 (6-12 m)

G

ro

u

n

d

w

a

te

r

ta

b

le

(

G

L

-m

)

Figure 4. Temporal change of the groundwater potential observed

at well DRK-W1 in deep aquifer (lower sand formation), and at

wells DRK-W2 and W3 in shallow aquifer (upper sand formation)

in 2005. The screens of observation wells at Darikandi site (Figure

Fig. 4, Nakaya et al.

(37)

0.2

0.15

0.1

0.05

0

Jan.

Mar

.

May

.

Jul.

Sep.

Nov

.

Feb.

Apr

.

Jun.

Aug.

Oct.

Dec.

Evapo-transpiration (m/mon.)

Time

Figure 5. Monthly evapo-transpiration estimated from the

average monthly temperature between 1957 and 1987 at the

Rangamati meteorological observatory (22.63

o

N, 92.20

o

E, 63 m

above sea level (ASL)), Bangladesh, using the Thornthwaite

equation (Thornthwaite, 1948).

Fig. 5, Nakaya et al.

(38)

3 4 5 6 7 8 0 24 48 72 96 120 144 168 G ro u n d w a te r h ea d ( m ) Time (month) 3 4 5 6 7 8 180 204 228 252 276 300 324 348 G ro u n d w a te r h ea d ( m ) Time (month) 3 4 5 6 7 8 349 351 353 355 357 359 Z = 0m, -15m, -42m Irrigation + 4L/day/person H-Lower H-upper H H H G ro u n d w at er h ea d ( m ) Time (month) 3 4 5 6 7 8 349 351 353 355 357 359 Z = 0m, -15m, -42m Irrigation + 100L/day/person H-Lower H-upper H H H G ro u n d w at er h ea d ( m ) Time (month)

(a)

(b)

(c)

cal :W2 cal :W2 cal :W1 cal :W1 cal :W3 cal :W3

(d)

Figure 6. Numerically simulated groundwater head for the 1st to 15th year (a) and 16th to 30th years

(b) in 30-years simulation of wells DRK-W1, W2, and W3 for groundwater usage of both 4 and 100

L/(person*day). (c) and (d) show the variation of simulated groundwater head in the 30th year for the 4

L/(person*day) and 100 L/(person*day) groundwater usage levels, respectively, with observed values at

Hcal :W1 Hcal :W2 & W3

Fig. 6, Nakaya et al.

Figure 6

(39)

Groundwater table (m ASL)

6 7 8

23.7

23.71

23.72

90.58

90.59

90.6

90.61

2006 Oct.

N

(

o

)

E (

o

)

0 1000 2000 3000

3000

2000

1000

0

X (m)

Y (

m)

Observed

Figure 7. Contour map of unconfined groundwater head computed

7.0 m

7.0 m

Fig. 7, Nakaya et al.

(40)

90.59 90.6 90.61 2 3 .7 2 3 .7 1 2 3 .7 2 90.59 90.6 90.61 2 3 .7 2 3 .7 1 2 3 .7 2 90.59 90.6 90.61 2 3 .7 2 3 .7 1 2 3 .7 2 90.59 90.6 90.61 2 3 .7 2 3 .7 1 2 3 .7 2 90.59 90.6 90.61 2 3 .7 2 3 .7 1 2 3 .7 2 90.59 90.6 90.61 2 3 .7 2 3 .7 1 2 3 .7 2 90.59 90.6 90.61 2 3 .7 2 3 .7 1 2 3 .7 2 90.59 90.6 90.61 2 3 .7 2 3 .7 1 2 3 .7 2 90.59 90.6 90.61 2 3 .7 2 3 .7 1 2 3 .7 2 -500 0 500 1000 1500 2000 2500 3000 3500 -500 0 500 1000 1500 2000 2500 3000 3500 Groundwater flow in Sep. - 100L/day/person Vmax=1.30m/mon., Z=-10m Y ( m ) X (m) 90.59 90.6 90.61 2 3 .7 2 3 .7 1 2 3 .7 2 90.59 90.6 90.61 2 3 .7 2 3 .7 1 2 3 .7 2 90.59 90.6 90.61 2 3 .7 2 3 .7 1 2 3 .7 2 -500 0 500 1000 1500 2000 2500 3000 3500 -500 0 500 1000 1500 2000 2500 3000 3500 Groundwater flow in Jan. - 100L/day/person Vmax=3.00m/mon., Z=-10m Y (m ) X (m) -500 0 500 1000 1500 2000 2500 3000 3500 -500 0 500 1000 1500 2000 2500 3000 3500 Groundwater flow in Feb. - 100L/day/person Vmax=4.41m/mon., Z=-10m Y (m ) X (m) -500 0 500 1000 1500 2000 2500 3000 3500 -500 0 500 1000 1500 2000 2500 3000 3500 Groundwater flow in Mar. - 100L/day/person Vmax=4.46m/mon., Z=-10m Y (m ) X (m) -500 0 500 1000 1500 2000 2500 3000 3500 -500 0 500 1000 1500 2000 2500 3000 3500 Groundwater flow in Apr. - 100L/day/person Vmax=2.85m/mon., Z=-10m Y (m ) X (m) -500 0 500 1000 1500 2000 2500 3000 3500 -500 0 500 1000 1500 2000 2500 3000 3500 Groundwater flow in May. - 100L/day/person Vmax=1.57m/mon., Z=-10m Y (m ) X (m) -500 0 500 1000 1500 2000 2500 3000 3500 -500 0 500 1000 1500 2000 2500 3000 3500 Groundwater flow in Jun. - 100L/day/person Vmax=1.94m/mon., Z=-10m Y (m ) X (m) -500 0 500 1000 1500 2000 2500 3000 3500 -500 0 500 1000 1500 2000 2500 3000 3500 Groundwater flow in Jul. - 100L/day/person Vmax=1.30m/mon., Z=-10m Y (m ) X (m) -500 0 500 1000 1500 2000 2500 3000 3500 -500 0 500 1000 1500 2000 2500 3000 3500 Groundwater flow in Aug. - 100L/day/person Vmax=1.36m/mon., Z=-10m Y (m ) X (m) -500 0 500 1000 1500 2000 2500 3000 3500 -500 0 500 1000 1500 2000 2500 3000 3500 Groundwater flow in Oct. - 100L/day/person Vmax=2.19m/mon., Z=-10m Y (m ) X (m) -500 0 500 1000 1500 2000 2500 3000 3500 -500 0 500 1000 1500 2000 2500 3000 3500 Groundwater flow in Nov. - 100L/day/person Vmax=2.04m/mon., Z=-10m Y (m ) X (m) -500 0 500 1000 1500 2000 2500 3000 3500 -500 0 500 1000 1500 2000 2500 3000 3500 Groundwater flow in Dec. - 100L/day/person Vmax=2.20m/mon., Z=-10m Y (m ) X (m) (a) (a) (j) (b) (c) (i) (f) (k) (h) (g) (e) (d) (l)

Fig. 8, Nakaya et al. Figure 8

(41)

-500 0 500 1000 1500 2000 2500 3000 3500 -500 0 500 1000 1500 2000 2500 3000 3500

Groundwater flow in Feb. (After 30 y.) Vmax=4.42m/mon., Z=-10m Y ( m ) X (m) 90.59 90.6 90.61 2 3 .7 2 3 .7 1 2 3 .7 2

1

2

3

4

5

6

-30 -20 -10 0 100 5 10 15 20 25 30 1 2 3 4 5 6 Z ( m )

Model Age (year) (g) -40 -30 -20 -10 0 10 20 150 200 250 300 350 400 450 500 No.1 (2050, 450, z) Z (m ) Y (m) A B C D E F G

Model age (year) A : 2.5 B : 7.5 C : 12.4 D : 18.5 E : 28.0 F : 45.5 G : 49.6 2.5 years End point -40 -30 -20 -10 0 10 20 2300 2400 2500 2600 2700 2800 2900 3000 No.2 (2350, 850, z) Z (m ) X (m) AB C D E F G

2.5 years Model age (year) A : 3.4, E : 63.5 B : 9.4, F : 100.4 C : 14.5, G : > 116 D : 23.3 -40 -30 -20 -10 0 10 20 2500 2600 2700 2800 2900 3000 No.3 (2650, 1250, z) Z (m ) X (m)

Model age (year) A : 5.5, F : B : 16.5, G : C : 21.5 D : 39.5 E : 58.9 A B C D E F G 2.5 years -40 -30 -20 -10 0 10 20 1960 1980 2000 2020 2040 2060 No.4 (2050, 1450, z) Z (m ) X (m) A B C D E

Model age (year) A : 3.5, F : B : 11.5, G : C : 19.4 D : 27.6 E : 45.6 F G -40 -30 -20 -10 0 10 20 580 600 620 640 660 680 700 No.6 (600, 600, z) Z (m ) X (m) A B C D E F G

Model age (year) A : 2.5 B : 7.5 C : 14.4 D : 22.5 E : 32.5 F : G : -(a) (b) (c) (f) (e) (d) -40 -30 -20 -10 0 10 20 1000 1100 1200 1300 1400 1500 No.5 (1800, 1000, z) Z (m ) Y (m) A B C D E F G

Model age (year) A : 7.5, F : 62.6 B : 16.5, G : C : 21.5 D : 30.5 E : 44.5

Fig. 9, Nakaya et al. Figure 9

(42)

0 500 1000 1500 2000 2500 3000 y ( m ) -30 -20 -10 0 10 0 500 1000 1500 2000 2500 3000 z ( m ) x (m) 0 500 1000 1500 2000 2500 3000 -30 -20 -10 0 10 z (m) -30 -20 -10 0 10 0 500 1000 1500 2000 2500 3000 z (m ) x (m) 0 500 1000 1500 2000 2500 3000 y ( m ) 0 500 1000 1500 2000 2500 3000 -30 -20 -10 0 10 z (m) -30 -20 -10 0 10 1800 1900 2000 2100 2200 2300 z ( m ) x (m) 300 500 700 900 1100 1300 1500 -30 -20 -10 0 10 z ( m ) y (m) As (µg/L) 700-1200 500-700 300-500 100-300 0-100 As (µg/L) 700-1200 500-700 300-500 (b) (a) (a) (b)

Figure 10. Simulated groundwater flow paths toward the midpoint of the 5-meter long screen, along with As concentrations for the 126 surveyed tube wells installed in the shallow (Holocene) aquifer (above about -20m ASL). (a) As <300 µg/L, (b) As >300 µg/L.

Figure 11. Vertical profile of simulated groundwater flow Fig. 11, Nakaya et al.

Fig. 10, Nakaya et al. Figure 10 and 11

(43)

500 m

Harihardi

Mamuddi

Darikandi

Temdi

Ledamdi

Fig. 12, Nakaya et al.

(44)

0

100

200

300

400

500

600

700

800

0

10

20

30

F

low

pa

th l

e

ngt

h (m

)

Model age (year)

As (µg/L)

700-1200

500-700

300-500

100-300

0-100

Figure 13. Relationship between the average length of the

groundwater flow path (Lp) and the average residence time

(model age) (TR) from the ground surface to tube wells, with the

As concentration of each well's water, excluding three wells that

had residence times >30 years (920 m, 38.5 years, 351 As µg/L),

V=8 m/y

V=34 m/y

Fig. 13, Nakaya et al.

References

Related documents

• In the event the polling location must be relocated to a different location, the City Clerk along with the Police Department will coordinate the move with assistance from

George is a substitute for the cellular calling experience that uses a caller’s already paid-for data plan instead of their local minutes.. Users can download the app, connect with,

In case where an employee retires during the year, the withholding agent conducts year-end tax settlement for the attributable year of retirement in obligation and issue the

Pending agreement to such a process RailCorp design staff shall monitor changes in operational control systems or Safeworking rules to identify changes that are likely to impact

In our opinion, the financial statements and the report of the board of Directors give a true and fair view of both the consolidated and the parent company’s financial performance

Clearly, software and database systems are needed that can handle large amounts of streaming time-series data in real time and that can also support analysis and management of

Therefore, the objectives of this work were to: (1) demonstrate that HMDS treatment can be used to increase strength of recycled corrugated fiberboard, and (2) show

 How do you implement Section Access on hierarchy based data.. QlikView Server &amp;