Retention and Release of Graphene Oxide in Structured

Sep 2, 2016 - In this work, saturated and unsaturated structured heterogeneous sand columns were used to examine the fate of graphene oxide (GO) nanop...
0 downloads 10 Views 813KB Size
Subscriber access provided by CORNELL UNIVERSITY LIBRARY

Article

Retention and Release of Graphene Oxide in Structured Heterogeneous Porous Media under Saturated and Unsaturated Conditions Shunan Dong, Xiaoqing Shi, Bin Gao, Jianfeng Wu, Yuanyuan Sun, Hongyan Guo, Hongxia Xu, and Jichun Wu Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.6b01948 • Publication Date (Web): 02 Sep 2016 Downloaded from http://pubs.acs.org on September 2, 2016

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Environmental Science & Technology is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 31

Environmental Science & Technology

1

Retention and Release of Graphene Oxide in Structured Heterogeneous Porous Media

2

under Saturated and Unsaturated Conditions

3

Shunan Dong1, Xiaoqing Shi1, Bin Gao3, Jianfeng Wu1, Yuanyuan Sun1*, Hongyan Guo2, Hongxia Xu1,

4

Jichun Wu1*

5 6

1

Key Laboratory of Surficial Geochemisty, Ministry of Education, School of Earth Sciences

7

and Engineering, Hydrosciences Department, Nanjing University, Nanjing 210023, China

8

2

9

Nanjing University, Nanjing 210093, China,

State Key Laboratory of Pollution Control and Resource Reuse, School of Environment,

10

3

Department of Agricultural and Biological Engineering, University of Florida, Gainesville,

11

FL 32611

12 13 14 15 16 17 18 19 20 21 22

__________________________

23

*

24

E-mail address: [email protected] (Y. Sun), [email protected] (J. Wu).

Corresponding authors. Tel.: +86 25 89680835; fax: +86 25 8368 6016.

25 1 ACS Paragon Plus Environment

Environmental Science & Technology

26



ABSTRACT ART Schematic

Transport

Fast Flow Domain

Slow Flow Domain

Saturated

Exchange

Slow Flow Domain

Fast Flow Domain

Unsaturated

Exchange

Release

Coarse Sand (700-850 µm) Fine Sand (450-500 µm)

27 28



29

In this work, saturated and unsaturated structured heterogeneous sand columns were used to

30

examine the fate of graphene oxide (GO) nanoparticles in heterogeneous porous media under

31

various conditions. A two-domain model considering mass exchange between zones was

32

applied to describe GO retention and transport in structured, heterogeneous porous media,

33

which matched the transport experimental breakthroughs well. Experimental and model

34

results showed that GO retention and transport in all the heterogeneous columns were

35

dominated by the preferential flow phenomena. Under saturated conditions, the coarse sand

36

with higher hydraulic conductivity was the fast-flow domain (FFD) and the fine sand was the

37

slow-flow domain (SFD), and both FFD and SFD affect GO particles fate in structured

38

heterogeneous media. When the heterogeneous columns were drained, the fine sand with

39

higher moisture content became the FFD and the coarse sand was the SFD, however,

40

preferential flows in the FFD dominated GO retention and transport processes. For all the

41

columns, the mobility of GO decreased with the increasing ionic strength (IS), and the

42

previous retained particles were released by reducing solution IS, indicating part of the

43

retained particles were trapped in the secondary minimum energy well.

ABSTRACT:

44 45 2 ACS Paragon Plus Environment

Page 2 of 31

Page 3 of 31

46

Environmental Science & Technology



INTRODUCTION

47

As a new carbon nanomaterial with exceptional physical and chemical properties,

48

graphene oxide (GO) has received increasing attention in many fields of applications,1,2 such

49

as nanoelectronics, conductive thin films, supercapacitors, nanosensors, and nanomedicine.3-9

50

Given the wide applications and rapid growth in production, it is expected that GO will

51

inevitably be released into the environment, even in groundwater systems with its relatively

52

high mobility in soils.10 Previous studies have already shown that GO nanoparticles are toxic

53

to organism and can reduce the activity of cells.11, 12 Understanding the transport behaviors of

54

GO in porous media is thus critical to evaluate the environmental impact and the potential

55

risk of this new nanomaterial.

56

A number of laboratory studies have been reported that the mobility of GO in porous

57

media is controlled by several subsurface environmental factors, including solution chemistry

58

(e.g., ionic strength and pH), flow rate, moisture content, particle concentration and surface

59

property, and media characters.13-22 These studies demonstrate that GO has the higher

60

mobility with lower ionic strength, higher flow rate, and the presence of natural organic

61

matter or anionic surfactant, meanwhile the transport of GO also may be limited by multiple

62

mechanisms, such as the decreasing media grain size, moisture content and input particle

63

concentration, and the presence of biofilms. In addition, it has been reported that GO can

64

serve as an effective carrier to facilitate the transport of polycyclic aromatic hydrocarbons in

65

saturated soils.23 Findings from previous studies also suggested that established theories and

66

models of colloid/nanoparticle transport in porous media can be applied to describe the

67

transport behaviors of GO particles.10, 13-16, 18-21 However, all of the previous studies of GO

68

fate and transport were conducted in homogenous porous media, which may not represent the

69

complex natural soil and groundwater systems well.24 3 ACS Paragon Plus Environment

Environmental Science & Technology

70

In fact, preferential flow is widely accepted from field and lysimeter observations as the

71

rule rather than the exception in various types of soils because of the natural heterogeneities.

72

25, 26

73

preferential flow phenomena and demonstrated that region with larger hydraulic conductivity

74

provides a fast path for solute transport (fast-flow domain) that dominates the mass transfer in

75

the heterogeneous porous media.24 Similarly, packed column experiments have shown that

76

the present of preferred flow paths within structured media may dominate colloid

77

transmission even though they occupy a small fraction of the porous media.27-29 Recently

78

studies also found that heterogeneities strongly affect the fate and transport of

79

microorganisms and engineered nanoparticles in porous media under different solution

80

chemistry conditions.30-32 These observational studies have been complemented by the

81

development of mathematical models using dual-permeability concepts or stochastic

82

approaches to describe preferential flow and the fate and transport of solute and particles in

83

heterogeneous porous media.30, 33-35 To the best of the authors’ knowledge, however, none of

84

the previous studies have systematically investigated the effect of structured heterogeneities

85

on the fate and transport of GO nanoparticles in porous media, which extremely limits the

86

prediction and monitoring on the fate of GO in the environment. Therefore, additional

87

investigations are still needed to promote the knowledge of retention and transport process

88

for GO in structured heterogeneous porous media.

89

Laboratory experiments with packed sand columns have also been used to examine the

The overarching objective of this work is to understand the effect of physical

90

heterogeneities on the transport and release of GO nanoparticles in saturated and unsaturated

91

porous media. Structured heterogeneous porous media was created by packing sand of two

92

different grain sizes in laboratory columns and GO transport behaviors were examined under

93

conditions of different moisture content and IS combinations. The specific objectives were as 4 ACS Paragon Plus Environment

Page 4 of 31

Page 5 of 31

Environmental Science & Technology

94

follows: (1) determine the coupled effects of physical heterogeneity and solution chemistry

95

(IS) on the retention and transport of GO in saturated and unsaturated porous media; (2)

96

examine the release (reentrainment) of previously retained GO as a result of IS perturbations

97

in the structured heterogeneous porous media; and (3) model GO retention and transport in

98

structured heterogeneous pore media under conditions of different moisture content and IS

99

combinations.

100 101 102



MATERIALS AND METHODS GO Suspension. Single-layer GO was purchased from ACS Material (Medford, MA).

103

As reported by the manufacturer, the lateral diameter of the GO is 1-5 µm and the layer

104

thickness is 0.8-1.2 nm. The actual physical dimensions of the GO have been determined

105

with atomic force microscopy (AFM) in a previous study with the average thickness of 0.92 ±

106

0.13 nm and average square root of the area of 582 ± 112 nm.36

107

To prepare the stock solution, 250 mg of GO were added into 1000 mL deionized (DI)

108

water and the mixture was then sonicated for 2 h for thorough dispersion. Prior to each

109

experiment, the stock solution was diluted to 25 mg L-1 with DI water and electrolyte solution.

110

Relatively high levels of GO were used in this work to better elucidate the interactions

111

between GO and porous media under various conditions. This practice is commonly used in

112

the literature and has successful advanced the knowledge of the environmental fate and

113

transport of GO.16, 17, 22 NaCl was used as the background electrolyte at two different

114

concentrations (1 and 20 mM). The pH for all the experimental solutions was adjusted to 5.6

115

with 1.0 mM HCl or NaOH. Only a small amount of HCl or NaOH solution was used in pH

116

adjustment, which had negligible effect on the solution IS. The GO suspensions prepared

117

under the experimental conditions were very stable.21, 36 5 ACS Paragon Plus Environment

Environmental Science & Technology

118

Concentration of the GO in the solution was determined by measuring the total

119

absorbance of light at a wavelength of 230 nm with a UV-2000 Spectrophotometer (UNICO

120

Instrument Co., Ltd. China) followed the method of Liu et al.18 The zeta potential values and

121

average hydrodynamic diameters of GO particles under varying solution chemistry conditions

122

were determined using a ZetaPALS (Brookhaven Instruments Corporation, NY, USA), and

123

the average hydrodynamic diameters of the particles were 474.2 ± 20 and 577.4 ± 24 nm for

124

1 mM and 20 mM IS conditions, respectively.

125 126

Sand. Quartz sand purchased from Unimin Corporation (Unimin-Le Sueur, MN, USA)

127

was used to make the experimental porous media. The sand was sieved into the different size

128

ranges: 700-850, 450-500, 350-450, 250-350, 150-200, and 110-150 µm. They were all

129

washed sequentially by tap water, 10% nitric acid (v: v) and deionized water, following the

130

procedures of Tian et al.37 The zeta potentials of the sand under different solution chemistry

131

conditions were measured following the method developed by Johnson et al.38 Briefly,

132

colloidal quartz sand suspensions were obtained by ultrasonication of sand in different

133

solution chemistry conditions for 10 min, and their zeta potential values were then measured

134

by the ZetaPALS. The surface morphological features of the sand were characterized with a

135

Scanning Electron Microscope equipped with Energy Dispersive X-ray (SEM-EDX, JEOL

136

JSM-6490, Japan).

137 138

Column Experiments. Vertical acrylic columns (20 cm long and 2.5 cm inside diameter

139

for saturated experiments, 16.7 cm long and 2.6 cm inside diameter for unsaturated

140

experiments) were used in the transport experiments. Stainless steel membranes with 50 µm

141

pores were used at both inlet (bottom) and outlet (top) to seal the columns and to distribute 6 ACS Paragon Plus Environment

Page 6 of 31

Page 7 of 31

Environmental Science & Technology

142

the flow. For the unsaturated column, six vent holes were drilled on opposite sides at 3, 7.5,

143

and 12 cm from the top of the column. The vent holes were sealed with gas-permeable porous

144

PTFE membranes (Milliseal Disk, Millipore) to allow air to enter under unsaturated

145

conditions. Solutions were pumped through the column at a constant rate of 2.0 mL min-1

146

using the peristaltic pump (BT100-1F, Longer Pump, Hebei, China). Relatively high flow rate

147

used in this work to better elucidate the transport process of GO in structured heterogeneous

148

porous media under various conditions, which is a common practice in literature to study the

149

fate and transport of nanoparticles and other contaminants.17, 39-41 All the transport

150

experiments were performed in duplicate.

151

Columns with a single structured, heterogeneity were constructed by placing a

152

thin-walled acrylic plate (thickness < 1 mm) in the center of the column to separate it into

153

two identical regions. To make a saturated, heterogeneous porous media column, sand of two

154

different size classes was carefully wet-packed into the two regions of the column using the

155

procedure similar to Wu et al.24 After the column was packed, the acrylic plate was then

156

slowly withdrawn. Six types of saturated, structured heterogeneous columns (labeled as

157

S1-S6) packed with different types of sands were used in the experiments. Table 1 lists

158

arrangement of the columns and the basic properties of the sand.

159

To make an unsaturated column, two peristaltic pumps (BT100-1F, Longer Pump, Hebei,

160

China) connected at the column inlet and outlet were used first to drain the column and then

161

to maintain steady-state flow using the procedure similar to Liu et al.18 Initially, a saturated

162

column was drained by elevating the outflow rate 5% higher than the inflow rate. When the

163

target moisture content was reached, the inflow and outflow rates were equalized to 2.0 mL

164

min-1. The average moisture content of the entire unsaturated structured columns was about

7 ACS Paragon Plus Environment

Environmental Science & Technology

165

0.20. Three types of unsaturated, heterogeneous columns (U1-U3) packed with different

166

types of sand combinations were used in the experiments (Table 1).

167

Each of the packed sand column was first flushed with ~4 pore volumes (PVs) of DI

168

water and then ~4 PVs of particle-free electrolyte solution (1 or 20 mM NaCl). A

169

breakthrough experiment was then initiated by injecting a ~2 PV pulse of the GO suspension,

170

followed by several PVs of particle-free background solution to flush out unretained GO in

171

porewater. For selected columns (i.e., IS of 20 mM ones), after flushing with the background

172

solution, DI water was applied to them to mobilize the previously deposited particles.

173

Effluent samples were collected using a fraction collector (BS-100A, Puyang Scientific

174

Instrument Research Institute, China). The concentrations of GO were measured immediately

175

after sample collection with the methods mentioned previously.

176

A tracer solution with 9.75 mM NaCl and 0.25 mM KNO3 (NO3− as the tracer) was

177

injected to the pre-equilibrated columns and the concentrations of nitrate in the effluent was

178

determined at a wavelength of 220 nm using a UV-2000 Spectrophotometer (UNICO

179

Instrument Co., Ltd. China).15 Additional tracer experiments for saturated and unsaturated

180

homogeneous sand columns (packed uniformly with each type of the sand) were also

181

conducted to determine the longitudinal dispersivity of GO for all the tested grain size and

182

moisture conditions (values are listed in Table 1).

183 184

XDLVO Theory. The extended DLVO (XDLVO) theory was used to quantify the

185

interaction energy between GO and sand grains, GO and GO, and GO and air-water interface.

186

The theory considers the effects of van der Waals attraction, electrical double layer repulsion

187

and Lewis acid-base interactions.15, 18, 36 Details of the XDLVO theory can be found in the

188

Supporting Information S1. 8 ACS Paragon Plus Environment

Page 8 of 31

Page 9 of 31

Environmental Science & Technology

189 190

GO Retention and Transport Model. The two-domain conceptualization42 was applied

191

in this work to simulate GO retention and transport in the columns, which are usually

192

described with two advection-dispersion equations:24, 30

193

∂ C FFD ∂ 2 C FFD ∂ C FFD ∂ C ' FFD + = D FFD − v FFD − αβ 2 ∂t ∂t ∂z ∂z

194

∂ C SFD ∂ 2 C SFD ∂ C SFD ∂ C ' SFD + = D SFD − v SFD + αβ 2 ∂t ∂t ∂z ∂z

FFD

SFD

( C FFD − C SFD ) (1)

( C FFD − C SFD )

(2)

195

where the subscripts FFD and SFD refer to the fast flow and slow flow domains in the

196

column, respectively; C is the GO concentration in pore water (M L-3); D is the dispersion

197

coefficient (L2 T-1); v is the velocity of pore water (L T-1); α is a first-order mass transfer

198

coefficient for GO exchange between the fast-flow and slow-flow domains (L T-1); β is a

199

geometry coefficient (L-1), which is described below; and C’ is the concentration of GO

200

retained in the porous media (M L-3). For unsaturated porous media, C’ accounts for GO

201

retention that takes place due to the presence of both air-water and solid-water interfaces.

202

The retention of GO in porous media can be described with various types of kinetics

203

expressions.15, 21, 43 To avoid over-parameterization, a second-order, irreversible kinetics

204

expression was used to describe the concentration of GO retained in two domains:

205

∂ C ' FFD C ' FFD = k FFD C FFD (1 − ) ∂t X FFD

(3)

206

∂ C ' SFD C ' SFD = k SFD C SFD (1 − ) ∂t X SFD

(4)

207

where k is a retention rate constant (T-1); X is the maximum retention capacity of the porous

208

media (M L-3). When the retention capacity greatly exceeds the retained GO concentration

9 ACS Paragon Plus Environment

Environmental Science & Technology

Page 10 of 31

209

(i.e., X >> C’), equations (3) and (4) reduce to first-order kinetic expressions, in which the

210

retention rate varies linearly with the pore water GO concentration (C).

211

The first-order mass transfer coefficient, α, is an empirical parameter that quantifies the

212

rate of colloid exchange between the adjacent domains. Gerke and van Genuchten44

213

suggested that the magnitude of this parameter depends on the size and shape of the soil

214

matrix and on the permeabilities at the interfaces of the two domains. The geometry

215

coefficient, β, reflects the ratio of total contact surface area between the two domains to the

216

total amount of water within a specific domain (FFD or SFD), can be written as:

217

β FFD =

218

β SFD =

S V FFD θ FFD

S V SFD θ SFD

(5)

(6)

219

where S is the contact area between coarse and fine grains packed in the column (L2); V is the

220

volume of each domain (L3); and θ is the moisture content.

221

The governing equations of this two-domain model (i.e., (1)-(6)) were solved

222

numerically with zero initial concentrations, a pulse-input boundary condition at the column

223

inlet, and a zero-concentration-gradient boundary condition at the outlet. The simulated GO

224

concentrations in effluents were calculated with:30

225 226

C =

C SFD v SFD V SFD θ SFD + C FFD v FFD V FFD θ FFD v SFD V SFD θ SFD + v FFD V FFD θ FFD

(7)

The Gauss-Levenberg-Marquardt algorithm implemented in a gradient-based, model

227

independent, auto-optimization program PEST45 was used to optimize the value of the model

228

parameters that minimized the sum-of-the-squared differences between model-calculated and

229

measured breakthrough concentrations. This model optimization method was first applied to

230

the tracer breakthrough-curve data to obtain the best-fit values of the first-order mass transfer 10 ACS Paragon Plus Environment

Page 11 of 31

Environmental Science & Technology

231

coefficient (α) for the heterogeneous columns prior to initiating the model simulation of GO

232

for optimized retention parameters. The first-order mass transfer rate (α) of GO particles were

233

assumed to be the same as those of the tracer, and GO concentrations in the effluents were

234

then simulated to determine the best-fit values of k and X. The pore water velocity (v),

235

dispersion coefficient (D) and moisture content (θ) of FFD and SFD were determined through

236

experimental measurements and model calculations, which were detailed in Supporting

237

Information S2.

238 239 240



RESULTS AND DISCUSSION XDLVO Results. The zeta potential values of GO and the six types of sand were

241

consistently negative under all the tested experimental conditions (Table S1, Supporting

242

Information), which were consistent with the results of previous studies that both GO and

243

quartz sand are negatively charged under normal conditions.18 The zeta potentials of GO and

244

sand at low IS (1 mM) were all lower than those at high IS (20 mM), which can be attributed

245

to the increase of IS compressed the electrical double layer.

246

The XDLVO interaction energy profiles between GO and GO showed energy barriers

247

around 1.0 × 106 kT µm-2 at both IS conditions and a secondary minimum of -152 kT µm-2 at

248

high IS (Figure S2, Supporting Information). These results confirmed that the GO

249

suspensions were stable under the experimental conditions, particularly when the IS was low.

250

The presence of secondary minimum at the high IS might cause tentative aggregation of GO

251

particles in the long run. Nevertheless, previous studies have demonstrated that GO

252

suspensions are stable enough for the period of column transport studies under similar or

253

even higher IS conditions.15, 18, 21

11 ACS Paragon Plus Environment

Environmental Science & Technology

254

The XDLVO energy profiles between GO and sand also showed strong primary energy

255

barriers (> 2.7 × 106 kT µm-2) under the two IS conditions (Figure S1, Supporting

256

Information), suggesting the experimental conditions were unfavorable for GO deposition

257

onto the sand grains in porous media. At the high IS (i.e., 20 mM), all the XDLVO curves

258

showed secondary minimum wells with depths ranged between -161 and -168 kT µm-2 (Table

259

S1, Supporting Information). In this work, GO particles thus may be retained in porous media

260

through deposition into the secondary minima under the high IS conditions. The retention,

261

however, is reversible if the secondary minimum well diminish during IS reduction.46

262

The XDLVO energy profiles between GO and air-water interface showed much lower

263

primary energy barriers (< 4.8 × 104 kT µm-2) and no sign of secondary energy minimum

264

under both IS conditions. Previous study with bubble column experiments has demonstrated

265

that GO particles do not attach to the air-water interfaces,18 which is consistent with the

266

XDLVO results in this work.

267 268

GO Retention and Transport in Saturated Heterogeneous Porous Media. All the

269

breakthrough curves (BTCs) were plotted as normalized effluent concentration (C/C0) as a

270

function of cumulative volume of flow through the column. Under saturated conditions, the

271

BTCs of GO under 1 mM had similar shapes with those of tracers in the six columns, which

272

demonstrated the two-peak (S2, S3, and S6) or two-stage (S1, S4, and S5) transport behaviors

273

depending on grain size arrangements (Figure 1), reflecting the typical preferential flow

274

phenomena.24 In general, the two-peak phenomena appeared when the two types of sand in

275

the packed columns had relatively large size differences, corresponding to large differences in

276

hydraulic conductivities (Table 1). In the S2, S3 and S6 columns (Figure 1b,c, and f,

277

KFFD/KSFD > 5.5), the two separated BTCs of GO under 1 mM corresponded to the quick 12 ACS Paragon Plus Environment

Page 12 of 31

Page 13 of 31

Environmental Science & Technology

278

(preferential) breakthrough from the fast flow domain (FFD, i.e., coarse-grained matrix) and

279

the slow breakthrough from the slow flow domain (SFD, i.e., fine-grained matrix),

280

respectively. In S1, S4 and S5 (Figure 1a, d, and e, KFFD/KSFD = 1.7~3.3) columns, the BTCs

281

showed a different transport behavior: (1) a period of rapidly increasing concentrations,

282

reflecting the delivery of GO through the FFD flow path, (2) a brief period of slowly

283

increasing concentrations, suggesting the fully breakthrough of the GO in the FFD and the

284

transmission of GO from SFD to FFD, and (3) a second period of rapid concentration

285

increasing, reflecting the delivery of GO through the SFD flow path in addition to the fully

286

breakthrough in the FFD. The descending limbs of the BTCs also exhibited this stair-step

287

pattern, presumably due to the flushing of the coarse-grained FFD flow path with

288

particle-free electrolyte solution, followed by displacement of the particle-free electrolyte

289

solution from the fine-grained SFD. These results demonstrate that heterogeneity in hydraulic

290

conductivity is one of the dominant physical factors governing the preferential flow and

291

transport behavior of both tracer and GO in structured heterogeneous porous media. The

292

BTCs of GO in the columns at 1 mM conditions were only slightly lower than those of the

293

tracers (Figure 1), indicating low retention of the particles in the saturated heterogeneous

294

porous media. Because the experimental conditions were unfavorable for GO deposition onto

295

the sand grains, more than 90% of the GO were recovered in the effluents after the flushing

296

with background solution for all the columns (Table S3, Supporting Information).

297

When the solution IS was high (20 mM), there were more retention of GO in the

298

saturated heterogeneous columns and the BTCs were much lower than those of the tracer and

299

GO at low IS (Figure 1). Mass balance calculations indicated that the recovery of GO in the

300

effluent of all the saturated heterogeneous columns dropped dramatically to 31.0%-57.2%

301

(Table S3, Supporting Information). The reduced mobility of the GO in columns was due to 13 ACS Paragon Plus Environment

Environmental Science & Technology

302

the deposition of particles onto the sand surfaces in secondary minimum wells, as indicated

303

by the XDLVO calculations (Figure S1, Supporting Information). SEM-EDX analysis of the

304

sand excavated from the column inlet at the end of the transport experiments confirmed the

305

attachment of the particles onto the grain surfaces (Figure S4, Supporting Information). The

306

importance of secondary minimum deposition to GO fate and transport in sand porous media

307

has been demonstrated in several previous studies.15, 18, 21 It is worth noting that increase of IS

308

almost eliminated GO transport out from the slow-flow dominants in all the six structured

309

heterogeneous columns (Figure 1), which might be attributed to several reasons: 1) the

310

retention of GO significantly increased with the decreasing of grain sizes at 20 mM;21 2)

311

compared to the FFD, only small amount of GO entered the SFD, increasing the retention of

312

GO in SFD;21, 47 and 3) in comparison with the FFD, the resident times of GO in the SFD

313

were long due to the low velocity, increasing the transmission of GO from SFD to FFD.

314

The two-domain model reproduced these breakthrough characteristics of all the

315

saturated experiments closely (Figure 1), with computation of R2 exceeding 0.94 (Table 2).

316

The model parameters of the retention and transport of the tracer and GO are listed in Table

317

2. The flow velocity (v) varied from 0.177 to 1.95 cm min-1 and the dispersion coefficients

318

(D) varied by a factor ranging from 0.0044 to 0.253 cm2 min-1. For all of the saturated

319

treatments, v values in the FFD were all greater than those in the SFD, confirming the

320

preferential flow phenomena. Similarly, all the D values in FFD also greater than those in the

321

SFD, which is consistent with previous observations that dispersion varies proportionately

322

with flow velocity.48 Calculations of the Peclet number ( Pe =

323

the column) ranged from 1.54 × 102 to 8.06 × 102 (Table S2, Supporting Information),

324

indicating that advection processes dominated the transport processes in both FFD and SFD

325

in the columns. The best-fit values of the mass-transfer coefficient (α) were sensitive to the

vH , where H is the height of D

14 ACS Paragon Plus Environment

Page 14 of 31

Page 15 of 31

Environmental Science & Technology

326

experimental conditions and decreased with the grain sizes in the SFD (Table 2). Calculations

327

of the Damkohler number for mass transfer ( Da MT =

328

(Table S2, Supporting Information). These computations revealed that inter-domain exchange

329

contributed to the mass transfer process in the saturated heterogeneous columns, as the time

330

scale for this mass exchange was comparable to that for advective transport in the columns.

331

βα H v

) ranged from 0.05 to 0.84

The simulated BTCs of GO were computed with the values of α obtained from the tracer

332

studies without adjustment. At 1 mM conditions, the deposition of GO onto the sand grains

333

was little and the deposition kinetics thus were described with the first-order expression. For

334

the same grain sizes, the best-fit k values of FFD were almost the same and ranged from

335

0.00214 to 0.00227 min-1 and 0.00416 to 0.00421 min-1 for S1-S3 and S4-S6, respectively

336

(Table 2). The best-fit k values of SFD (ranged from 0.00331 to 0.00514 min-1 for S1-S3 and

337

from 0.00486 to 0.00559 min-1 for S4-S6, Table 2) were increased with the decreasing grain

338

sizes, suggesting that, under the same chemistry conditions, grain size mainly affected GO

339

retention and k values.21 Consequently, for all the saturated columns, kFFD was always less

340

than the kSFD for all of the saturated treatments. The Damkohler numbers for GO deposition

341

( Da CD =

342

the time scales for particles deposition were relatively large and, for some cases, might not

343

have a pronounced effect on transporting through the saturated columns. This result is

344

consistent with the mass recovery calculations and the predictions of the XDLVO theory.

345

kH ) ranged from 0.0219 to 0.625 (Table S2, Supporting Information), indicating v

When the solution IS was 20 mM, the best-fit kSFD values were much larger than those at

346

IS of 1 mM (Table 2). Similar to that under 1 mM conditions, the best-fit k of FFD also very

347

similar and ranged from 0.0613 to 0.0655 min-1 for S1-S3 and from 0.130 to 0.134 min-1 for

348

S4-S6 under 20 mM (Table 2). In SFD, k increased and ranged from 0.179 to 0.225 min-1 and 15 ACS Paragon Plus Environment

Environmental Science & Technology

349

0.187 to 0.313 min-1 for S1-S3 and S4-S6 with the decreasing grain sizes, respectively (Table

350

2). All the k values under 20 mM IS were higher than the corresponding ones under 1 mM,

351

which were consistent with the previous results.15, 18 The Damkohler numbers for GO

352

deposition at 20 mM ranged from 0.63 to 35.0 with most of were higher than 1 (Table S2,

353

Supporting Information). This calculation indicated the deposition process played an

354

important role in controlling the fate and transport of GO in the columns under the tested

355

conditions, confirming the importance of secondary minima to the fate and transport of GO in

356

structured heterogamous porous media. The model-estimated maximum retention capacities

357

X for all domains in the saturated heterogeneous columns ranged from 34.4 to 49.8 mg L-1

358

with the XSFD were also all larger than XFFD (Table 2), which is similar to the values of

359

previous studies.21

360

The model simulations also provide the opportunity to quantify the mass recovery rates

361

for each domain of the structured heterogeneous columns (Table S3, Supporting

362

Information). The results indicated that most of the particles in effluents were from the FFD,

363

further indicating the dominance of preferential flow to the fate and transport of nanoparticles

364

in saturated heterogeneous porous media.

365 366

GO Retention and Transport in the Unsaturated Heterogeneous Porous Media. The

367

BTCs of the tracer and GO in the unsaturated heterogeneous columns (U1, U2, and U3) did

368

not show the two-peak or two-stage behaviors (Figure 2). The moisture content of the two

369

area in unsaturated columns were obtained from the simulation results of draining process

370

(Supporting Information S2), which were both very low (0.067-0.121) (Table 1) in the coarse

371

sand area within the three unsaturated heterogeneous columns, suggesting limited flow in the

372

coarse sand during the transport experiments. For this reason, the fine sand structure became 16 ACS Paragon Plus Environment

Page 16 of 31

Page 17 of 31

Environmental Science & Technology

373

the FFD and dominated the flow and mass transfer process; while the coarse sand was the

374

SFD. As the retained water barely remained in the SFD, almost all GO were transported

375

through FFD by the water current, demonstrated the dominance of preferential flow in

376

structured heterogeneous porous media under unsaturated conditions. This also explained

377

why the BTCs in unsaturated heterogeneous porous media were similar to the typical BTCs

378

observed in homogenous porous media.27

379

Consistent with the results of saturated columns, GO also showed higher retention in

380

the columns at 20 mM than that at 1 mM conditions. Mass balance calculations showed that

381

89.4%-90.6% and 46.6%-56.2% of GO were collected from the outlet at 1 and 20 mM IS,

382

respectively (Table S3, Supporting Information). Although the moisture of the FFD in

383

column U1 was lower than those in U2 and U3, the mass recovery rate of U1 was similar (IS

384

= 1 and 20 mM), indicating no/little deposition of GO onto the air-water interface. This is

385

consistent with the XDLVO calculations and the findings of previous studies.18

386

The two-domain model reproduced the GO breakthrough characteristics in the

387

unsaturated heterogeneous columns fairly well (Figure 2), with R2 larger than 0.94 (Table 2).

388

The best-fit values of the flow velocity in FFD were much greater than those in SFD for all

389

treatments with the ratio higher than 44 (Table 2). The contrast was biggest for column U2

390

(~65), where the flow velocity in the SFD was barely 0.036 cm min-1 (Table 2). These model

391

results also confirmed that fine-grained matrix changed to the FFD and dominated the flow in

392

the unsaturated heterogeneous columns. As the fitted velocity profiles for SFD was negligible

393

compared with the FFD, the model was simplified by neglecting the dispersion processes in

394

SFD when describing tracer and GO transported in all the unsaturated treatments. This

395

simplification did not reduce the accuracy of the model predictions of tracer transport in the

396

unsaturated columns. The dispersion coefficients (D) from unsaturated homogeneous sand 17 ACS Paragon Plus Environment

Environmental Science & Technology

397

columns (Supporting Information S2) for the FFD under the tested conditions varied from

398

0.099 to 0.291 cm2 min-1 (Table 2). Calculations of the Peclet number ranged from 1.57 × 102

399

to 4.15 × 102 (Table S2, Supporting Information), indicating the dominance of advection

400

process. The Damkohler number for mass transfer in the FFD and SFD ranged from

401

0.30-0.55 and 54.2-83.0 (Table S2, Supporting Information), respectively, indicating the

402

importance of the exchange process to the mass transfer in the two domains, especially the

403

SFD.

404

Similarly, the values of v, and D used in tracer study were same as in simulating GO

405

BTCs in unsaturated heterogeneous porous media (Figure 2). The model-estimated values of

406

the retention rate constants (k) in FFD were increased from 0.00405 to 0.00420 min-1 at IS of

407

1 mM and from 0.136 to 0.209 min-1 at 20 mM with the grain size decreased (Table 2). The

408

Damkohler numbers for GO particle deposition in the unsaturated columns ranged from

409

0.0247 to 15.6 under 1mM condition and 0.83 to 84.0 under 20mM condition (Table S2,

410

Supporting Information), suggesting that particles retention rates were comparable to their

411

transport speeds in the unsaturated columns. Again, because GO deposition in the unsaturated

412

columns at IS of 1 mM was very low, the first-order kinetics were used in the simulations. At

413

the IS of 20 mM, the model-estimated maximum retention capacities X ranged from 36.1 to

414

66.1 mg L-1 (Table 2), within the ranges of the values reported in previous studies.21

415 416

Release of Previously Retained GO. Under saturated conditions, reduction in IS

417

remobilized the previously retained GO in the six structured heterogeneous columns and all

418

the release curves showed two peaks (Figure 3), indicating GO release from both FFD and

419

SFD. The peak release concentrations of GO from the FFD (12.4 mg L-1-22.2 mg L-1), which

420

was higher than those from the SFD (1.3 mg L-1-13.2 mg L-1). Mass balance calculation also 18 ACS Paragon Plus Environment

Page 18 of 31

Page 19 of 31

Environmental Science & Technology

421

showed that FFD (14.1%-17.8%) released more GO particles than SFD (5.1%-14.7%) (Table

422

S4, Supporting Information), probably because more particles were retained in the

423

preferential paths during the deposition experiments. In addition, mass balance calculation

424

also showed that reduction of solution IS released 19.2%-32.0% of the total GO applied to

425

the columns, corresponding to 42.8%-69.0% of the previously retained particles (Table S4,

426

Supporting Information). This result indicated the IS perturbation remobilized most of the

427

particles retained on the sand grains in the saturated heterogeneous porous media. Previous

428

studies have demonstrated that the retention of colloids and nanoparticles in porous media is

429

reversible if the attachment of particles on the media is through secondary minimum.37, 49 In

430

this work, both the XDLVO calculations and column experimental data, particularly the

431

release BTCs, suggested that secondary minimum governed the retention and release of GO

432

particles in structured heterogeneous porous media. Several previous studies have also

433

demonstrated the importance of secondary minimum to GO fate and transport in homogenous

434

porous media under similar conditions.15, 17

435

Because most of the flow processes were in the FFD under unsaturated conditions, the

436

release curves of the three unsaturated columns only showed one peak (Figure 4), consistent

437

with the breakthrough curves. In comparison to those of the corresponding saturated

438

experiments, the peak release concentrations of the three unsaturated columns were slightly

439

lower and ranged from 7.6 mg L-1 to 13.5 mg L-1. Mass balance calculation showed that the

440

unsaturated columns (12.1%-22.8%) released less GO particles than the corresponding

441

saturated columns (21.1%-29.8%) (Table S4, Supporting Information). Furthermore, mass

442

balance calculation also showed that reduction of solution IS only released 27.6%-42.7% of

443

the previously retained GO particles in unsaturated columns. This result confirmed that, in

19 ACS Paragon Plus Environment

Environmental Science & Technology

444

addition to deposition onto grain surface, other mechanisms might contribute the retention of

445

GO in structured heterogeneous porous media under unsaturated conditions.18

446 447



ENVIRONMENTAL IMPLICATIONS

448

Findings from this work indicated that structured heterogeneities have strong effects on

449

GO retention, release, and transport in the porous media. Because structured heterogeneities

450

caused by natural and anthropological activities, are ubiquitous in the subsurface environment,

451

accurate characterizations of media structures thus are crucial to the prediction and

452

monitoring of the fate and transport of GO and other nanoparticles in soil and groundwater

453

systems. In particular, preferential flow arising from structured heterogeneities may dominate

454

the flow and mass transfer processes in both vadose zone and the groundwater, which may

455

cause the quick delivery of GO into aquifers and thus impose potential risks to public health.

456

The two-domain model was applied in this work and described the retention and transport of

457

GO in the structured, heterogeneous porous media very well. Nevertheless, further

458

investigations are needed to develop, refine, and optimize mathematical models based on the

459

two-domain conceptualizations, and thus to expand current capacities to better understand

460

and predict the fate and transport of GO in the subsurface.

461 462 463



ASSOCIATED CONTENT Details of XDLVO theory and relevant results, the determination of pore water velocity,

464

dispersion coefficient and moisture content, summary of Peclet and Damkohler numbers,

465

characterization of GO and sand and other additional information is available in the

466

supporting information.

467 20 ACS Paragon Plus Environment

Page 20 of 31

Page 21 of 31

468

Environmental Science & Technology



469

ACKNOWLEDGMENTS This work was supported by the National Natural Science Foundation of

470

China-Xianjiang project (U1503282), and the National Natural Science Foundation of China

471

(41372234, 41172207 and 41172206).

472 473 474



REFERENCE (1) Dikin, D. A.; Stankovich, S.; Zimney, E. J.; Piner, R. D.; Dommett, G. H.;

475

Evmenenko, G.; Nguyen, S. T.; Ruoff, R. S. Preparation and characterization of graphene

476

oxide paper. Nature 2007, 448 (7152), 457-460.

477

(2) Konkena, B.; Vasudevan, S. Understanding aqueous dispersibility of graphene oxide

478

and reduced graphene oxide through pKa measurements. J. Phys. Chem. Lett. 2012, 3 (7),

479

867-872.

480

(3) Dong, S.; Sun, Y.; Wu, J.; Wu, B.; Creamer, A. E.; Gao, B. Graphene oxide as filter

481

media to remove levofloxacin and lead from aqueous solution. Chemosphere 2016, 150,

482

759-764.

483

(4) Zhang, C.; Chen, S.; Alvarez, P. J. J.; Chen, W. Reduced graphene oxide enhances

484

horseradish peroxidase stability by serving as radical scavenger and redox mediator. Carbon

485

2015, 94, 531-538.

486

(5) Liu, M.; Zhao, H.; Chen, S.; Yu, H.; Quan, X. Colloidal graphene as a transducer in

487

homogeneous fluorescence-based immunosensor for rapid and sensitive analysis of

488

microcystin-LR. Environ. Sci. Technol. 2012, 46 (22), 12567-12574.

489 490

(6) Allen, M. J.; Tung, V. C.; Kaner, R. B. Honeycomb carbon: A review of graphene. Chem. Rev. 2009, 110 (1), 132-145.

21 ACS Paragon Plus Environment

Environmental Science & Technology

491

(7) Duch, M. C.; Budinger, G. S.; Liang, Y. T.; Soberanes, S.; Urich, D.; Chiarella, S. E.;

492

Campochiaro, L. A.; Gonzalez, A.; Chandel, N. S.; Hersam, M. C. Minimizing oxidation and

493

stable nanoscale dispersion improves the biocompatibility of graphene in the lung. Nano Lett.

494

2011, 11 (12), 5201-5207.

495

(8) Jiang, Y.; Wang, W.; Liu, D.; Nie, Y.; Li, W.; Wu, J.; Zhang, F.; Biswas, P.; Fortner, J.

496

D. Engineered crumpled graphene oxide nanocomposite membrane assemblies for advanced

497

water treatment processes. Environ. Sci. Technol. 2015, 49 (11), 6846-6854.

498

(9) Peralta-Videa, J. R.; Zhao, L.; Lopez-Moreno, M. L.; de la Rosa, G.; Hong, J.;

499

Gardea-Torresdey, J. L. Nanomaterials and the environment: A review for the biennium

500

2008-2010. J. Hazard. Mater. 2011, 186 (1), 1-15.

501 502 503

(10) Qi, Z.; Zhang, L.; Chen, W. Transport of graphene oxide nanoparticles in saturated sandy soil. Environ. Sci.: Processes Impacts 2014, 16 (10), 2268-2277. (11) Chen, Y.; Ren, C.; Ouyang, S.; Hu, X.; Zhou, Q. Mitigation in multiple effects of

504

graphene oxide toxicity in zebrafish embryogenesis driven by humic acid. Environ. Sci.

505

Technol. 2015, 49 (16), 10147-10154.

506

(12) Zhao, J.; Wang, Z.; White, J. C.; Xing, B. Graphene in the aquatic environment:

507

Adsorption, dispersion, toxicity and transformation. Environ. Sci. Technol. 2014, 48 (17),

508

9995-10009.

509

(13) Fan, W.; Jiang, X.; Lu, Y.; Huo, M.; Lin, S.; Geng, Z. Effects of surfactants on

510

graphene oxide nanoparticles transport in saturated porous media. J. Environ. Sci. 2015, 35,

511

12-19.

512

(14) Fan, W.; Jiang, X.; Yang, W.; Geng, Z.; Huo, M.; Liu, Z.; Zhou, H. Transport of

513

graphene oxide in saturated porous media: Effect of cation composition in mixed Na-Ca

514

electrolyte systems. Sci. Total Environ. 2015, 511, 509-515. 22 ACS Paragon Plus Environment

Page 22 of 31

Page 23 of 31

515 516

Environmental Science & Technology

(15) Feriancikova, L.; Xu, S. Deposition and remobilization of graphene oxide within saturated sand packs. J. Hazard. Mater. 2012, 235, 194-200.

517

(16) He, J.; Li, C.; Wang, D.; Zhou, D. Biofilms and extracellular polymeric substances

518

mediate the transport of graphene oxide nanoparticles in saturated porous media. J. Hazard.

519

Mater. 2015, 300, 467-474.

520

(17) Lanphere, J. D.; Luth, C. J.; Walker, S. L. Effects of solution chemistry on the

521

transport of graphene oxide in saturated porous media. Environ. Sci. Technol. 2013, 47 (9),

522

4255-4261.

523

(18) Liu, L.; Gao, B.; Wu, L.; Morales, V. L.; Yang, L.; Zhou, Z.; Wang, H. Deposition

524

and transport of graphene oxide in saturated and unsaturated porous media. Chem. Eng. J.

525

2013, 229, 444-449.

526

(19) Liu, L.; Gao, B.; Wu, L.; Sun, Y.; Zhou, Z. Effects of surfactant type and

527

concentration on graphene retention and transport in saturated porous media. Chem. Eng. J.

528

2015, 262, 1187-1191.

529

(20) Qi, Z. C.; Zhang, L. L.; Wang, F.; Hou, L.; Chen, W. Factors controlling transport of

530

graphene oxide nanoparticles in saturated sand columns. Environ. Toxicol. Chem. 2014, 33

531

(5), 998-1004.

532

(21) Sun, Y.; Gao, B.; Bradford, S.; Wu, L.; Chen, H.; Shi, X.; Wu, J. Transport,

533

retention, and size perturbation of graphene oxide in saturated porous media: Effects of input

534

concentration and grain size. Water Res. 2015, 68, 24-33.

535

(22) Xia, T.; Fortner, J. D.; Zhu, D.; Qi, Z.; Chen, W. Transport of sulfide-reduced

536

graphene oxide in saturated quartz sand: Cation-dependent retention mechanisms. Environ.

537

Sci. Technol. 2015, 49 (19), 11468-11475.

23 ACS Paragon Plus Environment

Environmental Science & Technology

538

(23) Qi, Z.; Hou, L.; Zhu, D.; Ji, R.; Chen, W. Enhanced transport of phenanthrene and

539

1-naphthol by colloidal graphene oxide nanoparticles in saturated soil. Environ. Sci. Technol.

540

2014, 48 (17), 10136-10144.

541

(24) Wu, L.; Gao, B.; Tian, Y.; Munoz-Carpena, R. Analytical and experimental analysis

542

of solute transport in heterogeneous porous media. J. Environ. Sci. Health, Part A:

543

Toxic/Hazard. Subst. Environ. Eng. 2014, 49 (3), 338-343.

544

(25) Baveye, P.; Boast, C. W.; Ogawa, S.; Parlange, J. Y.; Steenhuis, T. Influence of

545

image resolution and thresholding on the apparent mass fractal characteristics of preferential

546

flow patterns in field soils. Water Resour. Res. 1998, 34 (11), 2783-2796.

547

(26) Morales, V. L.; Parlange, J. Y.; Steenhuis, T. S. Are preferential flow paths

548

perpetuated by microbial activity in the soil matrix? A review. J. Hydrol. 2010, 393 (1-2),

549

29-36.

550 551 552

(27) Mishurov, M.; Yakirevich, A.; Weisbrod, N. Colloid transport in a heterogeneous partially saturated sand column. Environ. Sci. Technol. 2008, 42 (4), 1066-1071. (28) Schelde, K.; Moldrup, P.; Jacobsen, O. H.; De Jonge, H.; De Jonge, L. W.;

553

Komatsu, T. Diffusion-limited mobilization and transport of natural colloids in macroporous

554

soil. Vadose Zone J. 2002, 1 (1), 125-136.

555

(29) Bradford, S. A.; Bettahar, M.; Simunek, J.; Van Genuchten, M. T. Straining and

556

attachment of colloids in physically heterogeneous porous media. Vadose Zone J. 2004, 3 (2),

557

384-394.

558

(30) Lv, X.; Gao, B.; Sun, Y.; Dong, S.; Wu, J.; Jiang, B.; Shi, X. Effects of grain size

559

and structural heterogeneity on the transport and retention of nano-TiO2 in saturated porous

560

media. Sci. Total Environ. 2016, 563, 987-995.

24 ACS Paragon Plus Environment

Page 24 of 31

Page 25 of 31

561

Environmental Science & Technology

(31) Wang, Y.; Bradford, S. A.; Simunek, J. Physicochemical factors influencing the

562

preferential transport of escherichia coli in soils. Vadose Zone J. 2014, 13 (1); DOI:

563

10.2136/vzj2013.07.0120.

564

(32) Wang, Y.; Bradford, S. A.; Simunek, J. Transport and fate of microorganisms in

565

soils with preferential flow under different solution chemistry conditions. Water Resour. Res.

566

2013, 49 (5), 2424-2436.

567 568 569

(33) Leij, F. J.; Bradford, S. A. Colloid transport in dual-permeability media. J. Contam. Hydrol. 2013, 150, 65-76.

(34) Bekhit, H. M.; Hassan, A. E. Stochastic modeling of colloid-contaminant transport

570

in physically and geochemically heterogeneous porous media. Water Resour. Res. 2005, 41

571

(2), W02010.

572

(35) Lassabatere, L.; Yilmaz, D.; Peyrard, X.; Peyneau, P. E.; Lenoir, T.; Simunek, J.;

573

Angulo-Jaramillo, R. New analytical model for cumulative infiltration into dual-permeability

574

soils. Vadose Zone J. 2014, 13 (12); DOI: 10.2136/vzj2013.10.0181.

575

(36) Wu, L.; Liu, L.; Gao, B.; Muñoz-Carpena, R.; Zhang, M.; Chen, H.; Zhou, Z.;

576

Wang, H. Aggregation kinetics of graphene oxides in aqueous solutions: Experiments,

577

mechanisms, and modeling. Langmuir 2013, 29 (49), 15174-15181.

578

(37) Tian, Y.; Gao, B.; Silvera-Batista, C.; Ziegler, K. J. Transport of engineered

579

nanoparticles in saturated porous media. J. Nanopart. Res. 2010, 12 (7), 2371-2380.

580

(38) Johnson, P. R.; Sun, N.; Elimelech, M. Colloid transport in geochemically

581

heterogeneous porous media: Modeling and measurements. Environ. Sci. Technol. 1996, 30

582

(11), 3284-3293.

25 ACS Paragon Plus Environment

Environmental Science & Technology

583

(39) Yang, X.; Zhang, Y.; Chen, F.; Yang, Y. Interplay of natural organic matter with

584

flow rate and particle size on colloid transport: Experimentation, visualization, and modeling.

585

Environ. Sci. Technol. 2015, 49 (22), 13385-13393.

586

(40) Jiemvarangkul, P.; Zhang, W.; Lien, H. Enhanced transport of polyelectrolyte

587

stabilized nanoscale zero-valent iron (nZVI) in porous media. Chem. Eng. J. 2011, 170 (2-3),

588

482-491.

589

(41) Gardea-Torresdey, J. L.; Contreras, C.; de la Rosa, G.; Peralta-Videa, J. R. Flow rate

590

and interference studies for copper binding to a silica-immobilized humin polymer matrix:

591

Column and batch experiments. Bioinorg. Chem. Appl. 2005, 3 (1-2), 1-14.

592 593 594

(42) Skopp, J.; Gardner, W. R.; Tyler, E. J. Solute movement in structured soils-2-region model with small interaction. Soil Sci. Soc. Am. J. 1981, 45 (5), 837-842. (43) Bradford, S. A.; Torkzaban, S.; Leij, F.; Simunek, J. Equilibrium and kinetic models

595

for colloid release under transient solution chemistry conditions. J. Contam. Hydrol. 2015,

596

181, 141-152.

597

(44) Gerke, H.; van Genuchten, M. T. A dual-porosity model for simulating the

598

preferential movement of water and solutes in structured porous media. Water Resour. Res.

599

1993, 29 (2), 305-319.

600 601 602

(45) Doherty J., Brebber L., Whyte P. PEST, Model-Independent Parameter Estimation-User Manual; Watermark Computing: Corinda, Australia, 1994.

(46) Hahn, M. W.; O'Melia, C. R. Deposition and reentrainment of Brownian particles in

603

porous media under unfavorable chemical conditions: Some concepts and applications.

604

Environ. Sci. Technol. 2004, 38 (1), 210-220.

26 ACS Paragon Plus Environment

Page 26 of 31

Page 27 of 31

605

Environmental Science & Technology

(47) Bradford, S. A.; Kim, H. N.; Haznedaroglu, B. Z.; Torkzaban, S.; Walker, S. L.

606

Coupled factors influencing concentration-dependent colloid transport and retention in

607

saturated porous media. Environ. Sci. Technol. 2009, 43, (18), 6996-7002.

608 609 610

(48) Marino, M. A. Longitudinal dispersion in saturated porous-media. J. Hydraul. Div., Am. Soc. Civ. Eng. 1974, 100 (NHY1), 151-156.

(49) Shen, C.; Li, B.; Huang, Y.; Jin, Y. Kinetics of coupled primary-and

611

secondary-minimum deposition of colloids under unfavorable chemical conditions. Environ.

612

Sci. Technol. 2007, 41 (20), 6976-6982.

613 614 615 616 617 618 619 620 621 622 623 624 625 626 627 628 27 ACS Paragon Plus Environment

Environmental Science & Technology

629

Page 28 of 31

Table 1. Summary of Conditions for All Experiments

630 Saturated Column

Overall Pore Volume

dCS

dFS

θFFD

θSFD

KSFD

KFFD -1

(µm, FFD)

(µm, SFD)

700-850 700-850 700-850 350-450 350-450 350-450

450-500 250-350 150-200 250-350 150-200 110-150

0.354 0.354 0.354 0.354 0.354 0.354

0.344 0.357 0.364 0.357 0.364 0.384

dCS

dFS

θFFD

θSFD

(µm, SFD)

(µm, FFD)

(cm min )

(cm min-1)

19.8 19.8 19.8 6.0 6.0 6.0

7.8 3.6 1.8 3.6 1.8 0.6

pH

ω

5.6 5.6 5.6 5.6 5.6 5.6

0.53 0.55 0.54 0.54 0.55 0.58

pH

ω

(mL)

S1 S2 S3 S4 S5 S6 Unsaturated Column

38.3 37.9 38.6 38.1 39.2 39.5 Overall Pore Volume

KFFD

KSFD

(cm min-1)

(cm min-1)

(mL)

U1 33.6 700-850 450-500 0.292 0.121 --5.6 0.54 U2 33.5 700-850 250-350 0.346 0.067 --5.6 0.55 U3 34.3 350-450 150-200 0.315 0.094 --5.6 0.53 Where dCS, dFS, θ, K and ω represent diameter of coarse sand, diameter of fine sand, moisture content, saturated hydraulic conductivity, and the ratio of the volumes of the coarse sand area and the total sand system, respectively; FFD and SFD represent fast flow domain and slow flow domain. In this work, coarse sand and fine sand in saturated columns were FFD and SFD, respectively. In the unsaturated columns, however, coarse sand and fine sand became SFD and FFD, respectively. K were determined by the constant head method based on the Darcy’s Law. θ of FFD and SFD were determined indirectly, which were detailed in Supporting Information S2.

631 632

28 ACS Paragon Plus Environment

Page 29 of 31

633 634

Environmental Science & Technology

Table 2. Summary of Model Parameters for All Experiments

Column

vFFDa

vSFDa -1

DFFDa -1

2

-1

DSFDa 2

α* -1

-1

k FFD**

IS 1 mM kSFD**

2

R

k FFD**

kSFD**

IS 20 mM XFFD**

XSFD**

(min-1)

(min-1)

(mg L-1)

(mg L-1)

R2

(cm min )

(cm min )

(cm min )

(cm min )

(cm min )

(min-1)

(min-1)

S1

1.64

0.644

0.212

0.0615

0.00586

0.00227

0.00331

0.97

0.0655

0.179

34.4

38.2

0.96

S2

1.81

0.329

0.234

0.0168

0.00294

0.00218

0.00474

0.95

0.0649

0.188

36.1

39.6

0.96

S3

1.95

0.177

0.253

0.0069

0.00244

0.00214

0.00514

0.98

0.0613

0.225

38.0

39.9

0.98

S4

1.40

0.842

0.085

0.0431

0.00848

0.00421

0.00486

0.97

0.134

0.187

40.3

46.7

0.97

S5

1.65

0.496

0.100

0.0192

0.00409

0.00418

0.00508

0.94

0.133

0.231

41.5

48.0

0.97

S6

1.79

0.179

0.108

0.0044

0.00165

0.00416

0.00559

0.95

0.130

0.313

44.1

49.8

0.97

U1

2.74

0.057

0.291

--

0.0247

0.00405

0.0336

0.98

0.136

0.180

53.8

36.1

0.94

U2

2.35

0.036

0.132

--

0.0135

0.00417

0.0337

0.99

0.141

0.181

55.2

36.7

0.94

U3

2.46

0.055

0.099

--

0.0183

0.00420

0.0348

0.99

0.209

0.246

66.1

39.1

0.94

where v, D, α, k, and X represent velocity of pore water, the dispersion coefficient, the first-order mass transfer coefficient, retention rate constant, and the maximum retention capacity of the porous media, respectively; FFD and SFD represent fast flow domain and slow flow domain; superscript “a” means parameters were determined indirectly (Supporting Information S2); superscript “*” means parameters determined from the tracer experiments of Column S1 to S6 and Column U1 to U3; superscript “**” means parameters determined from the GO transport experiments. In this work, coarse sand and fine sand in saturated columns were FFD and SFD, respectively; in the unsaturated columns, however, coarse sand and fine sand became SFD and FFD, respectively.

29

ACS Paragon Plus Environment

Environmental Science & Technology

635

Figure Captions

636

Figure 1. Observed and fitted breakthrough curves of tracer and GO transport in saturated

637

heterogeneous pore media under various solution chemistry conditions (a: S1; b: S2; c: S3; d:

638

S4; e: S5; and f: S6).

639

Figure 2. Observed and fitted breakthrough curves of tracer and GO transport in unsaturated

640

heterogeneous pore media under various solution chemistry conditions (a: U1; b: U2; and c:

641

U3).

642

Figure 3. Observed breakthrough curves of GO during the release process in saturated

643

heterogeneous pore media (a: S1; b: S2; c: S3; d: S4; e: S5; and f: S6).

644

Figure 4. Observed breakthrough curves of GO during the release process in unsaturated

645

heterogeneous pore media (a: U1; b: U2; and c: U3).

646 647

Figure 1

648 649 650 30 ACS Paragon Plus Environment

Page 30 of 31

Page 31 of 31

651

Environmental Science & Technology

Figure 2 (a)

(b)

(c)

652 653 654

Figure 3

655 656

Figure 4 (a)

(c)

(b)

657

31 ACS Paragon Plus Environment